You are on page 1of 246

Compliant Leverage Mechanism Design for MEMS Applications

by Xiao-Ping Susan Su

B.S. (Tsinghua University) 1991 M.S. (University of California, Davis) 1997

A dissertation submitted in partial satisfaction of the requirements for the degree of Doctor of Philosophy

in Engineering-Mechanical Engineering in the GRADUATE DIVISION of the UNIVERSITY OF CALIFORNIA, BERKELEY

Committee in charge: Professor Alice M. Agogino, Chair Professor Dennis K. Lieu Professor Tsu-Jae King

Spring 2001

COMPLIANT LEVERAGE MECHANISM DESIGN FOR MEMS APPLICATIONS

Copyright 2001 by Xiao-Ping Susan Su

ABSTRACT
Compliant Leverage Mechanism Design for MEMS Applications
by Xiao-Ping Susan Su Doctor of Philosophy in Mechanical Engineering University of California, Berkeley Professor Alice M. Agogino, Chair

Compliant microleverage mechanisms, including single- and multistage, can be used in micro-electro-mechanical system (MEMS) to transfer an input force/displacement to an output to achieve mechanical and/or geometry advantages. This thesis presents the original systematical study on this mechanism with primary focus on the design theory and synthesis issues of the mechanism. Starting from the basic nomenclature, definition and classification of the mechanism, an extensive first-order analytical model and a second-order refined one are built for the single-stage microleverage mechanism, the basic element of all microleverage mechanisms. The amplification factor depends not only on the ideal leverage ratio (L/l), the geometry of the lever, but also the axial and bending spring constant of the output system. Good agreement is obtained between the results of secondorder analytical modeling and those of FEM simulation with SUGAR, a MEMS simulation tool. The compliance match theory developed applies to the two-stage microleverage mechanism. The axial spring constant of the lever stage close to the output needs to be in

a specific region (Region II and III) in order for the entire mechanism to effectively amplify force. The increased resistance of a microlever to rotational or axial displacement when the output and pivot are at the different side of the lever arm leads to lower amplification factor. The maximum amplification factor of a multistage microlever was derived in terms of the given output system and minimum flexure beam dimension. The design of microleverage mechanism in a resonant accelerometer is illustrated at each step to present the theory. The leverage mechanism for displacement amplification is analyzed with application in a disk-drive suspension and a micro-valve. Experimental verification of the analytical equations and SUGAR simulation was carried out at both the micro- and macro-scale. A 1S-2D (first stage first kind lever with output and pivot at same side, second stage second kind with pivot and output at different side) type of mechanism was fabricated by the SOI-MEMS process for inertial force amplification in a resonant accelerometer. A macro-scale aluminum model was built and the testing results agree qualitatively with the analytical and SUGAR predictions.

_______________________________ Professor Alice M. Agogino Chair of the Committee

ii

To My Family:

Dr. Henry S. Yang Jenny Su Yang Rachel Su Yang

iii

ACKNOWLEDGMENTS
I would like to extend my heartfelt thanks to my research advisor Prof. Alice Agogino for her encouragement, guidance, and advice for my research. Without her encouragement and support, this research would have been given up in many times and this thesis would never come out. Many thanks also go out to my thesis committee, Professors Agogino, Dennis Lieu, and Tsu-Jae King, for helping me put this document together. During the journey, Prof. Roger Howe, Prof. Dennis Lieu, Prof. Liwei Lin have given valuable insight to and constructive advice on the research and the dissertation writing. During my graduate study here at Berkeley, Prof. Hedrick, Prof. Pello, Prof. Kazerooni all have helped me in different avenues. This research initially followed the work by Dr. Trey Roessig. Thanks go to Dr. Timothy Brosnihan for processing the SOI resonant accelerometer, and Ms. Ningning Zhou for many discussions in SUGAR simulation. Ms. Jocelyn Lee showed me how to bond the chips. Many friends I met here at Berkeley have helped me in many ways. There are too many to mention here. Special thanks are due to my husband, Dr. Henry S. Yang, for building the aluminum model and a lot of tedious work in polishing the papers that came out of this research. My mother-in-law has helped taking care of my children during my Ph.D. study.

iv

I also would like to acknowledge the financial support of a GSR by Prof. Howe through a M3S DARPA contract 1-442427-25316(F30602-97-2-0266) (97-98), a NSF graduate research fellowship (96-99) and a Department GSI support (2000). To anyone who is interested in reading this thesis, I would like to share the great secret of my life, which is the Lord Jesus. My strength is from Him, who died and raised again. Everything I did is not on my own strength and my own ability, but by Him. He is faithful and has plans for each individual, no matter how common we are. And He can always uses our life, our brokenness to do something beautiful. Good and bad things are all from Him. He loves to work with us in difficult situations. And the Words said With Him, you can do anything. He said: In good time, I am with you. In bad times, I am carrying you through. and From glory to glory, I am changing you. The great secret of success, prosperity, happiness and joy in life is to love the Lord and to do His will.

TABLE OF CONTENTS
Abstract Dedication .................................................................................................................... iii Acknowledgments ....................................................................................... iv Table of Contents ....................................................................................................... vi List of Figures ............................................................................................................. xi List of Tables ............................................................................................................. xix Nomenclature ..............................................................................................................xx Chapter 1 Introduction .............................................................................. 1 Chapter 2 Compliant Microleverage Mechanism .............................. 7
2.1 2.2 2.3 2.4 2.5 Leverage Mechanism and Governing Laws ...............................................7 Compliant Leverage Mechanism ................................................................8 Classification of Compliant Leverage Mechanisms ...................................9 Design of Microleverage Mechanism and Its Challenge ..........................11 FEM Simulation ........................................................................................13

Chapter 3 Flexure Pivot Design .............................................................. 16


3.1 3.2 3.3 Flexure Pivot Design .................................................................................16 Different Models of Flexure Pivot ............................................................19 Pivot and Output System Configurations .................................................21

vi

Chapter 4 Single-stage Microleverage Mechanism ............................... 27


4.1 4.2 Basic Structure of a Single-stage Microleverage Mechanism ..................27 Amplification Factor of Single-Stage Microleverage Mechanism ...........30 4.2.1 4.2.2 4.3 4.4 Second-kind single-stage microleverage mechanism ...................30 First-kind microleverage mechanism ............................................34

Amplification Coefficient, A* ...................................................................35 Effect of Output System and Pivot Spring Constants on the Amplification Factor ................................................................................36

4.5 4.6 4.7 4.8 4.9

Second-order Refined Analytical Model ..................................................41 Spring Constant Calculation .....................................................................46 Strain Energy Analysis and Mechanical Efficiency .................................47 Mechanical Advantage and Geometry Advantage ...................................50 Single-Stage Microleverage Mechanism Design in a Resonant Output Accelerometer ...............................................................50 4.9.1 4.9.2 Resonant output accelerometer (RXL) .........................................50 Optimization of single-stage microleverage mechanism in the resonant accelerometer ....................................56

4.10

Other Design Issues of Microleverage Mechanism ..................................70 4.10.1 Beam column strength ..................................................................70 4.10.2 Effect of horizontal force ..............................................................71

Chapter 5 Two-stage Microleverage Mechanism ................................. 73


5.1 5.2 Two-Stage Microleverage Mechanism Structure .....................................74 Two-Stage Microleverage Mechanism Amplification Factor ..................76 vii

5.3 5.4 5.5

Compliance Relationship Between Adjacent Levers ...............................80 Different Configurations of Two-Stage Leverage Mechanism ................87 Two-Stage Microleverage Mechanism Design and optimization in the Resonant Accelerometer .....................................................................92

Chapter 6 Multiple-stage Microleverage Mechanism ........................109


6.1 6.2 Analysis of Multi-stage Microleverage Mechanism ...............................110 Maximum Amplification Factor of Multistage Lever for a Given Output System ..............................................................................113 6.3 Multi-stage Microleverage Mechanism Design in the Resonant Accelerometer ..............................................................116 6.3.1 6.3.2 Estimation of maximum achievable amplification factor ...........117 Feasibility analysis of a three-stage microleverage mechanism ..................................................................................118 6.4 Compliance in Microleverage Mechanism .............................................120

Chapter 7 SOI-MEMS Fabricated Resonant Accelerometer ............122


7.1 7.2 7.3 7.4 7.5 7.6 7.7 SOI-MEMS Process Run ........................................................................122 SOI-MEMS Fabricated Resonant Accelerometer ...................................124 Two-stage Microleverage Mechanism Optimization ..............................129 DETF Natural Frequency ........................................................................139 Sensitivity Analysis of the Resonant Accelerometer ..............................145 Optimization of The DETF Resonator ....................................................147 Testing of the SOI-MEMS Fabricated RXL ...........................................149

viii

7.7.1 7.7.2 7.7.3 7.8

Experiment Testing Set-ups.........................................................149 Testing of the DETF Resonator ..................................................150 Testing of the Accelerometer ......................................................152

Future Work on Resonant Accelerometer ...............................................154

Chapter 8 Experimental Verification with Macro Model ..................156


8.1 8.2 Building the Macro Aluminum Model ...................................................156 Experimental Verification with the Macro Model ..................................160

Chapter 9 Microleverage Mechanism for Displacement Amplification 165


9.1 Different Configurations of Leverage Mechanism for Displacement Amplification .........................................................................................164

9.2

Amplification Factor of Leverage Mechanism for Displacement Amplification ..........................................................................................169

9.3

Displacement Amplification Leverage Mechanism in a Silicon Disk-Drive Suspension ...............................................................................................172

9.4

Application of Leverage Mechanism for Displacement Amplification in a Silicon Microvalve...................................................................................180

Chapter 10 Conclusions and Contributions .......................................... 191

ix

Bibliography .............................................................................................. 195 Appendix


A. B. C. D.

................................................................................................. 202
ABAQUS Input File for Single-stage Microlever Simulation ................206 Netlist Files for Different Pivot Design .................................................209 Mathematica File For Analytical Analysis .............................................210 Netlist File for Single-Stage Microleverage Mechanism in Resonant Accelerometer .........................................................................................212

E.

Netlist File for Two-stage Microleverage Mechanism in Resonant Accelerometer .........................................................................................213

F. G. H.

Netlist File for Calculating Resonator Frequency ..................................214 Netlist File for Simulation of the DETF Resonator ................................215 Netlist File for Simulation of the 1S-2D type Two-stage Leverage Mechanism in the Macro-model .216

I.

Netlist File for Simulation of the 1S-2S type Two-stage Leverage Mechanism in the Macro-model .217

J.

Netlist File for Simulation of the Two-stage Leverage Mechanism in the Disk-drive suspension .218

LIST OF FIGURES Chapter 2


Fig. 2.1 Fig. 2.2 Mesh generated by ABAQUS for the leverage mechanism. Mechanism deflection simulated by ABAQUS.

Chapter 3
Fig. 3.1 Fig. 3.2 Fig. 3.3 Fig. 3.4 The general flexure hinge model. A flexure pivot fabricated by the SOI-MEMS technology. A flexure beam pivot fabricated by the SOI-MEMS technology. Different pivot models: (a) vertical pivot, (b) horizontal pivot, (c) combined pivot. Fig. 3.5 Comparison of the amplification factors of three models as a function of the pivot beam length Fig. 3.6 Fig. 3.7 Effect of pivot beam with different angles on the amplification factor A second-kind leverage mechanism with pivot and the output system (a) on the same side, and (b) at different sides of the lever arm. Fig. 3.8 Fig 3.9 The shape of a deflected pivot located on the opposite side of the lever arm. The deflection of pivot and connection beams of a 2D microlever.

xi

Chapter 4
Fig. 4.1 Schematic of three kinds of microleverage mechanisms (a) first kind, (b) second kind, (c) third kind. Fig. 4.2 (a) A second-kind microleverage mechanism before and after loading; (b) Model of the second-kind microlever under loading. Fig. 4.3 Amplification factor as a function of output system axial spring constant for a series of (a) pivot lengths and (b) pivot widths at a fixed output system rotational spring constant k m o of 2 x 10-7 Nm; (c) output system bending spring constant. Fig. 4.4 Amplification factor as a function of output system beanding spring constant for a series of output system axial spring constant. Fig. 4.5 Free-body diagram of a second kind micro-leverage mechanism under loading. Fig. 4.6 A layout of the resonant-output microaccelerometer with a proof-mass, two resonators, and four symmetrical single-stage second-kind microlevers. Fig. 4.7 (a) Schematic of DETF as the output system in the microaccelerometer; (b) simulated compound spring connecting seven springs. Fig. 4.8 The node information for SUGAR simulation of single-stage microleverage mechanism in the resonant accelerometer.

xii

Fig. 4.9

Effect of tuning fork beam width wf on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and at the pivot to the input energy.

Fig.4.10

Effect of T-F joint beam width wj on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and at the pivot to the input energy.

Fig. 4.11 Effect of pivot beam width wp on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and at the pivot to the input energy. Fig. 4.12 Amplification factor as a function of tuning-fork and joint beam width (tuning fork beam has the same width as that of joint beam) at various pivot beam width, 0.5 5 m. Fig. 4.13 Effect of T-F beam length lf on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and at the pivot to the input energy. Fig. 4.14 Effect of T-F joint beam length lj on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and at the pivot to the input energy.

xiii

Fig. 4.15 Effect of pivot beam length lp on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and at the pivot to the input energy. Fig. 4.16 The effect of lever ratio, L/l, on the amplification factor for a series of beam widths ( all beams are assumed to have same width). Fig. 4.17 The effect of the distance between pivot and tuning fork, l, on the amplification factor for a fixed level ratio L/l of 21. Fig. 4.18 The effect of lever arm width on the amplification factor by SUGAR simulation.

Chapter 5
Fig. 5.1 A schematic of a two-stage microleverage mechanism consisting of a firstkind lever as the first stage and a 2nd-kind as the second stage. Fig. 5.2 Regime classification of a typical plot of the amplification factor as a function of the axial spring constant. Fig. 5.3 Input axial spring constant of the overall first stage calculated by different methods. Fig. 5.4 Fig. 5.5 Comparison of the amplification factor of different configuration. (a) A 1S-1D type two-stage microleverage mechanism for RXL. (b) A 1S-2D type two-stage microleverage mechanism for RXL.

xiv

(c) A 1S-2D type two-stage microleverage mechanism for RXL. (d) A 1S-2S type two-stage microleverage mechanism for RXL. (e) A 2S-2D type two-stage microleverage mechanism for RXL. (f) A 2S-2S type two-stage microleverage mechanism for RXL. Fig. 5.6 (a) Amplification factors, A1, A2, and A; and (b) first-stage spring constant k and second-stage amplification factor as a function of the width of first-stage pivot, tuning fork and connection beam width (A series of beam width of the second stage were selected to see their effect on the second-stage amplification and total amplification factors). Fig. 5.7 Amplification factors, A1, A2, and A as a function of (a) T-F width; (b) T-F connection beam width; (c) lever 1 pivot width; (d) lever 1 pivot and connection width. Fig. 5.8 Amplification factors, A1, A2, and A as a function of (a) T-F beam length; (b) connection beam length; (c) first-stage lever pivot length. Fig. 5.9 Comparison of SUGAR result with the analytical result.

Fig. 5.10 Effect of (a) the second-stage connection width; (b) the second-stage pivot width on Amplification factors, A1, A2, and A. Fig. 5.11 Effect of (a) The second stage connection length; (b)The second stage pivot length on Amplification factors, A1, A2, and A.

xv

Chapter 7
Fig. 7.1 Fig. 7.2 Cross section schematic of the SOI-MEMS process run. (a) SEM of the resonant accelerometer with two-stage microleverage mechanism fabricated by SOI-MEMS run. (b) Layout Design of the Resonant Accelerometer Fig. 7.3 Fig. 7.4 SEM of the on-chip trans-resistance amplifier. Schematic of the trans-resistance amplifier circuit including (a) biasing supply, (b) amplifier, and (c) output stage. Fig. 7.5 Fig. 7.6 Fig. 7.7 SEM micrograph showing the connection of the lever to proofmass. SEM of the lateral anchor by the SOI-MEMS process. Schematic of a two-stage microleverage mechanism for the resonant accelerometer and layout for SOI integrated run. Fig. 7.8 Fig. 7.9 SEM of the two-stage microleverage mechanism fabricated by SOI-MEMS. The Amplificaiton factor as a function of (a) the width and (b) the length of the tuning fork. Fig. 7.10 The amplification factors as a function of (a) the width and (b) length of the first stage pivot. Fig. 7.11 The amplification factors as a function of (a) the width and (b) length of the connection beam 2 between the two lever stages.

xvi

Fig. 7.12 The amplification factors as a function of (a) the width and (b) length of the second stage pivot. Fig. 7.13 SEM of the SOI-MEMS fabricated DETF resonator. Fig. 7.14 SUGAR simulation of the DETF resonator with comb-drive mass. Fig. 7.15 SUGAR simulation of electrical tuning of DETF resonator frequency. Fig. 7.16 Sensitivity as a function of the DETF beam width. Fig. 7.17 Sensitivity as a function of the DETF beam length. Fig. 7.18 SOI-MEMS RXL Post-Process Steps. Fig. 7.19 Testing apparatus for the SOI-MEMS resonant accelerometer. Fig. 7.20 Off-chip unit gain buffer schematic. Fig. 7.21 Plots of magnitude and phase of the DETF resonator. Fig. 7.22 Schematic of a two-Axis resonant accelerometer with two-stage

Microleverage mechanism (a) 1S-2S, (b) 2S-2S.

Chapter 8
Fig. 8.1 Fig. 8.2 Fig. 8.3 Fig. 8.4 Aluminum backing plate for anchoring the leverage mechanism. Scaled-up two-stage aluminum leverage mechanism. A photograph of the macro aluminum leverage mechanism. Experimental set-up for macro-model verification. xvii

Fig. 8.5

Measured amplification factor as a function of the input load for two-stage aluminum microlevers: 1S-2S and 1S-2D.

Chapter 9

Fig. 9.1 Fig. 9.2

Two kinds of leverage mechanisms for displacement amplification. Test structure of two-stage microleverage mechanism for displacement amplification.

Fig. 9.3

(a) PZT-actuated silicon suspension (Chen and Horowitz). (b) Schematic of the silicon suspension design (Chen and Horowitz).

Fig. 9.4 Fig. 9.5 Fig. 9.6

Two-stage leverage mechanism design in the disk drive suspension. Node information of the two-stage mechanism in the disk drive suspension (a) The amplification factor as a function of the 2nd pivot width. (b) The amplification factor as a function of the 2nd pivot length.

Fig. 9.7

(a) The effect of 1st pivot width on the amplification factor. (b) The effect of 1st pivot width on the amplification factor.

Fig. 9.8 Fig. 9.9

Z-direction vibration mode displayed by SUGAR. Y-direction vibration mode displayed by SUGAR.

Fig. 9.10 (a) Schematic of a silicon microvalve with single-stage leverage mechanism (Williams 1999); (b) Two-stage microlever in the silicon microvalve.

xviii

Fig. 9.11 (a) Effect of first-stage pivot width change on the amplification factor. (b) Effect of first-stage pivot length change on the amplification factor. Fig. 9.12 (a) Effect of connection beam width change on the amplification factor. (b) Effect of connection beam length change on the amplification factor. Fig. 9.13 (a) Effect of second-stage pivot width change on the amplification factor. (b) Effect of second-stage pivot width change on the amplification factor. Fig. 9.14 A three-stage leverage mechanism in the silicon microvalve. Fig. 9.15 (a) Amplification factor change with the 1st pivot width. (b) Amplification factor change with the 1st pivot length. Fig. 9.16 (a) Amplification factor change with 2nd pivot width. (b) Amplification factor change with 2nd pivot length. Fig. 9.17 (a) Amplification factor change with the 3rd pivot width. (b) Amplification factor change with 3rd pivot length.

xix

List of Tables

Chapter 3

Table 3.1 Table 3.2

SUGAR Simulation Results of Different Pivot Models. Comparison of SUGAR results on Microlevers with Pivot and Output System on (a) the Same Side and (b) Different Sides of the Lever Arm.

Chapter 5

Table 5.1

Comparison of Different Configurations of Two-Stage Mechanisms in the resonant accelerometer

Table 5.2

Beam Deflections and Amplification Factor of Various Types of 2-stage Microleverage Mechanisms.

Chapter 9

Table 9.1

Different Configurations of Two-stage Microleverage Mechanism for Displacement Amplification.

xx

NOMENCLATURE
Ai A* Ama x E force amplification factor of the i -stage microlever
th

Amplification Coefficient maximum achievable force amplification factor of a compound microlever Youngs modulus of the microleverage mechanism material the mechanical efficiency of a microleverage mechanism, as defined by the ratio of output energy to input energy

ei

the efficiency of i

th

microlever stage, as defined by the ratio of output

energy to input energy for the stage Fin Fout Fi input force output force Axial force at the input of i microlever stage
th th th

kvvc, i

vertical (axial) spring constant of the beam connecting the (i-1) to the i microlever stage

kvvI, i

input vertical (axial) spring constant of the i microlever stage connected together with all the upstream lever stages and the output system

th

kvvo

the vertical (axial) spring constants of the external output system

xxi

kvvo, i kvvp, i k m c, i

vertical (axial) spring constant at the output of the i microlever stage


th

th

vertical (axial) spring constant of the pivot for the i microlever stage
th th

bending spring constant of the beam connecting the (i-1) microlever stage

to the i

k m I, i

input bending spring constant of the i

th

microlever stage connected

together with all the upstream lever stages and the output system

k m o k m o, i k m p, i
lc , i lp , i li

the bending spring constants of the external output system


th

output system bending spring constant of the i microlever stage


th

bending spring constant of the pivot of the i microlever stage


th

the length of connection beam of the i microlever stage


th

the length of pivot beam of the i microlever stage


th

length between the pivot and output for the i lever arm
th

Li wc , i wp , i
o

length between the input and output for the i lever arm
th

the width of connection beam of the i microlever stage


th

the width of pivot beam of the i microlever stage

vertical (axial) displacement at the external output system

xxii

I , i

vertical (axial) displacement at the input of i microlever stage

th

kh h o, khhp

the horizontal spring constants at output and pivot, respectively, when loaded with a horizontal force

kh m o, khmp

the horizontal spring constants at output and pivot, respectively, when loaded with a bending moment

kvvo, kvvp

the vertical (axial) spring constants of the output system and the pivot, respectively, when loaded with a vertical force

k h o, k h p

rotational spring constant of the output system and the pivot, respectively, when loaded with a horizontal force

k m o, k m p

rotational spring constant of the output system and the pivot, respectively, when loaded with a bending moment

Mo, Mp Uin Uvo, U o

bending moments at the output and pivot work performed by the input force (static loading) the tensile/compression strain energy and moment-bending strain energy, respectively, consumed at the output system

Uvp, U p

the tensile/compression strain energy and moment-bending strain energy, respectively, consumed at the pivot

tc , tf , tp

the thickness of all flexure beams and the lever arm

xxiii

mo, h o

the rotation angle of the lever arm under loading the rotation angle of the output caused by bending moment and horizontal force

mp, h p

the rotation angle of the pivot caused by bending moment and horizontal force

xxiv

CHAPTER 1 INTRODUCTION

MEMS is a rapidly-growing field which mainly involves the fabrication of microscale mechanical devices with integrated circuit process technology. Micro-motors as small as the diameter of a human hair have been fabricated with silicon and their functionality demonstrated. Many other micro-devices such as integrated sensors and actuators were reported to have been successfully fabricated with thin film deposition and etching methods. Machines are built at micro-scale, with components controlled by different actuation methods. Micro-mechanisms--including gears, four-bar linkages, and micro-robotics--could be all integrated together. MEMS devices ranging from automobile crash sensors (inertial sensors), to drug delivery systems (microfluidics), to hard-drive suspension arms will impact our everyday life in the very near future. MEMS also brings many challenges to the fundamental science of many disciplines because different prevalent forces and governing equations will come into play when the dimension changes from macro-scale to micro-scale. The length scale is a fundamental quantity that dictates the type of forces governing physical phenomena. With body forces scaled to the third power of the length scale, size effects can be seen in most physical phenomena at the micro-scale. With surface forces depending on the first or second power of the characteristic length, surface effects also become dominant in the micro-world. Many phenomena at the micro-scale are different from those at macro-scale. For example,
-1-

mixing of two different fluids at the micro-scale takes much longer since diffusion rates are different from that at the macro-scale. The frictional force is a function of the contact surface at the microscale instead of the traditional frictional law f = N . Deviations from conventional thinking and understanding are commonly encountered at the micro-world, creating new scientific frontiers as well as new technologies. Mechanisms that are well understood in the macro-world become the subject of research in the micro-scale. This thesis focuses on the systematic study of the lever in micro-scale design method and several design applications. Leverage mechanisms (Mach, 1960 and Ohanian, 1991), including single-stage and multi-stage, are very useful in MEMS in order to transfer an input to an output to achieve mechanical and/or geometrical advantage. Similar to its wide application at the macro-world, micro-leverage mechanisms can be used to transfer an input force/displacement to an output, either amplifying them or changing the force directions (Sniegowski, 1995; Keller, 1995; Lin, 1993; etc). The design of micro-leverage mechanisms is different from that of the conventional levers in the macro-world. With fabrication technology constraints, the microleverage mechanism is mainly formed by coplanar flexures, with one end of the flexure beam anchored to the substrate as a pseudopivot. Mechanical transformation in a microleverage mechanism is achieved by elastic deformation of its component flexure beams. This group of mechanism is the so-called compliant mechanism. Compliant mechanisms transmit force and/or motion through flexibility and undergo elastic deformation. Salanon (1989) defined a complaint mechanism as a mechanism that gains all or part of its mobility from the relative flexibility of its
-2-

members (Salamon 1989 and Howell 1991). Compliant mechanisms have many advantages compared with the conventional rigid-body mechanisms. They require fewer parts, no or minimal assembly, experience less wear and noise, less problems associated with friction and backlash. For applications in MEMS planar microstructures, compliant mechanisms provide a new group of micro-devices such as compliant four-bar linkages and micro-grippers. Compared with the earlier micro-mechanical structures such as cantilevers and diaphragms, this new group of micro-devices can perform more types of mechanical tasks than cantilevers and diaphragms could do, e.g., tools for microfabrication, nanofabrication, microsurgery and nanoprobing analysis systems. In microsurgery where accuracy is required in microscopic and very sensitive operations, the micro-mechanisms can function as precise tweezers and knives. In microanalysis and nanoanalysis systems, the micro-mechanisms can be used as positioning tools or for precision probing of a surface (Larsen, 1997). With the advancement of micro-fabrication technology, both compliant mechanisms and compliant microstructures can be designed to have desired material properties such as specific thermal expansion coefficients, negative Poissons ratio and piezoelectric behavior (Kikuchi 1998 and Larsen 1997). Compliant microleverage mechanisms can be widely used as mechanical amplifiers in micro-grippers, micro-actuators, micro-fluidics and micro-sensors, or wherever a force or displacement transformation is needed. There are several papers in the open literature on compliant microleverage mechanisms including both single-stage and multiple-stage compound microlevers designed to amplify either force or displacement (Sniegowski, 1995; Keller, 1995; Lin, 1993; etc). Snegowski and Smith (1995) presented a leverage mechanism that converts a

-3-

short-displacement, high-force electrostatic actuation to a long-displacement, mediumforce output. Keller (1995-1997) designed a lever structure with an amplification factor of 27 to transfer a thermal actuation to the motion of a micro-tweezer tip. Roessig (19951998) used leverage mechanisms to increase the sensitivity in a resonant accelerometer by amplifying the inertial force with both asymmetrical and symmetrical structures. Wad et al. (1996) at the Jet Propulsion Laboratory developed a two-stage mechanical flexure amplifier with pairs of piezoelectric stacks for vibration suppression. Other leverage mechanisms have been designed to amplify stroke of piezoelectric stacks (Main and Garcia, 1997). William (1999) used a mechanical lever in the operation of a micro-valve. The analysis and synthesis of compliant mechanisms can be very difficult and time consuming due to their kinematic complexity and deflection non-linearity. Analysis of these mechanisms is typically accomplished by finite element methods. Synthesis of compliant mechanisms, i.e., obtaining a suitable topology, shape and size of the mechanism for performing a specified mechanical operation, has caught the attention of many researchers (Kota et al. 1994). From the literature, there are two major approaches: the kinematics synthesis approach and the continuum synthesis approach. The kinematics synthesis approach is also called the pseudo-rigid-body model approach (Howell, 1994) where flexure mechanisms are designed by creating a kinematic model of the basic mechanism configuration, then replacing the mechanism hinge joint by flexure hinges. This method normally requires an initial mechanism geometry configuration. The continuum synthesis approach uses the topology optimization technique for mechanical structures. Ananthasuresh et al. (1994) presents a homogenization method where the structure optimization approach is used.

-4-

Although compliant leverage mechanisms have great potential application in MEMS, the theory and design issues are not well understood and documented. Designers usually take a trial-and-error, intuitive and iterative approach. Most of the devices are custom-designed for particular applications on an ad-hoc basis. The basic kinematic mechanism configurations are generated based on intuition and experience, or by replacing conventional joints in a rigid-link mechanism with flexure hinges. The flexure designs are then refined through a trial-and-error or finite element methods. Although good accuracy is required, the amplification factor of a leverage mechanism can only be roughly estimated in most cases. In this dissertation, the design of compliant microleverage mechanism are systematically studied and presented with a direct analytical method approach and finite element method approach, followed by both micro- and macro-scale experimental verifications and design applications. Chapter 2 introduces the basic structure, classification and nomenclature of microleverage mechanisms. Chapter 3 mainly discusses different flexure pivot designs in microleverage mechanisms. Analysis of a single-stage micro-leverage mechanism with a first-order model and a second-order refined model are presented in Chapter 4 to derive the equations for calculating the amplification factor. Related design issues are also discussed. An energy analysis and the concept of mechanical and geometrical advantages of a single-stage microleverage mechanism are also introduced. The design method introduced for a single-stage micro-leverage mechanism lays the foundation for not only the single-stage microleverage mechanism design, but also the two-stage and multiplestage microleverage mechanism design.
-5-

Based on the theories found for single-stage, the design of two-stage microleverage mechanism is presented in Chapter 5. Multiple-stage microleverage mechanism designs are discussed in Chapter 6. The maximum achievable amplification factor for a specific output system is calculated in terms of the output system spring constant and the lever ratio. The design of a microleverage mechanism in a resonantoutput micro-accelerometer is presented throughout this dissertation to illustrate the design issues and study the effect of lever ratio and the dimensions of pivot, connection beams, and tuning fork beams on the amplification factor. A resonant accelerometer with a two-stage microleverage mechanism is designed and fabricated with the SOI-MEMS processing. Chapter 7 focuses on several important issues concerning the SOI-MEMS fabricated resonant accelerometer. These issues include frequency and sensitivity calculations, the effect of thickness changes on the sensitivity and amplification factors of the mechanism, testing procedures and the analysis of experimental results. Chapter 8 presents the experimental verification of the design theory on a macro-scale aluminum model and the scaling-factor issues. Chapter 9 presents the design of microleverage mechanisms for displacement amplification with two application cases, one is for a diskdrive suspension and the other for a microvalve. Finally, the principal findings and the key contributions of this research work are summarized in Chapter 10.

-6-

CHAPTER 2 COMPLIANT MICROLEVERAGE MECHANISMS

In the World Book Encyclopedia A605, it is stated that the lever was first described by Greek scientist and philosopher Archimedes who once said, Give me a long enough lever and a place to stand, I can move the earth. Others claim that it is the Greek philosopher Aristotle who first described lever theory in his book titled Mechanics (Mach, 1960; Beckwith et al.). Regardlessly, all agree that the applications of levers can be found widely in industry and nature. While levers are found in automobiles, bikes, weight machines, etc., many parts in human body (e.g., arm and jaw) also function as leverage mechanisms. It is a simple machine making many job functions easier to perform, transferring an input to an output to achieve mechanical and/or geometrical advantages. It can change the force direction, e.g., from pushes to pulls, and also amplify force or displacement.

2.1 Leverage Mechanism and the Governing Laws


A mechanism is defined as a device for the coupling and transforming of energies (Hall, 1953). Alternatively, Sandor and Erdman (1984) define the mechanism as a mechanical device that has the purpose of transferring motion and/or force from a

-7-

source to an output. When a lever is used to transfer an input motion and/or force to an output one, it is called a leverage mechanism. The governing laws in a leverage mechanism are force and moment balance and energy conservation. In an ideal situation (a rigid lever arm without bending and a perfect pivot with free rotation or rigid support), the moment with respect to the pivot point should be balanced, Fin Lin = Fout Lout, while the lever is statically balanced. The mechanical work done by the input force should equal to that done by the output force, Fin in = Fout out if no strain energy is consumed at the pivot or by bending of any flexure beam. For a compliant mechanism the input energy should be equal to the output energy plus the elastic bending energies of the flexible components. When a lever is used to perform a function, a trade-off is made between the force and displacement. For example, to lift an object to a certain height, a greater force is needed to move a shorter distance while a smaller force is needed to travel a longer distance.

2.2 Compliant Leverage Mechanisms


Similar to its wide range of applications in the macro-world, levers have been widely used in microelectromechanical systems, and are called microleverage mechanisms. With current microfabrication technology constraints, a microleverage mechanism is typically formed by planar flexures, achieving mechanical transformation through the elastic bending of its members. This group of microleverage mechanisms is also called compliant microleverage mechanisms.

-8-

Compliant microleverage mechanisms can transfer an input (e.g., force or displacement) to an output to achieve mechanical advantage and/or geometrical advantage in MEMS, such as changing force directions between pushes and pulls and amplifying force or displacement. For example, in the case of an actuator a leverage mechanism can attenuate the force and the displacement generated by the original actuation mechanism to a desired output. Other applications include magnifying inertial forces in inertial sensors such as micro-accelerometers and gyroscopes to increase sensitivity, amplifying the tensile force in a tensile testing machine, and tuning a microresonator. Similar to the role of the operational amplifier (common-source amplifier and multi-stage amplifier) in microelectronics, microleverage mechanisms (single- and multiple-stage) are mechanical amplifiers in MEMS.

2.3 Classifications of Compliant Leverage Mechanisms


A compliant leverage mechanism consists of four major parts: lever arm (rigid part), pivot, the input and output systems. Depending on the different positions of the pivot, the input and the output system on the lever arm, the leverage mechanisms can be classified into three kinds. The first-kind is defined by the pivot lying between the input and the output. When the output lies in the middle of the pivot and the input, it is called a second-kind lever. When two lever arms of a second-kind lever join together at a hinge, it is called a double second-kind lever. A third-kind lever is defined by the input placed between the pivot and the output. The third-kind lever is used to amplify displacement. A second-kind lever is mainly for force amplification. A first-kind lever can amply either

-9-

force or displacement depending on the distance between the input and the pivot and the distance between the output and the pivot. When the arms of two third-kind levers join together as in the case of a pair of tweezers, the leverage mechanism is called a double third-kind lever. For each kind of compliant leverage mechanism, there are two subgroups: (i) one group with the output system and the pivot beam on the same side of a lever arm (designated as S group); and (ii) the other group with the output system and the pivot beam on the different sides of a lever arm (designed as D group). A single-stage leverage mechanism can be identified by a code starting with a number (1, 2 or 3) corresponding to the lever kind and followed by a letter (S or D) specifying the subgroup, e.g., 1S, 1D, 2S, 2D, 3S, 3D. A compound leverage mechanism or multiple-stage leverage mechanism is formed by stacking multiple stages of lever together. Different kinds of levers can be stacked together. If individual single-stage levers are properly configured, their amplification factors will be multiplied. In a compound leverage mechanism, each lever stage needs to be identified for reference. There are major two kinds of classifications: downstream classification (from output to input) and upstream classification (from input to output). In the downstream classification, the one connecting to the output system is called the first-stage, then the second-stage, until the input system. In the upstream classification, the one connected to the input system is called the first-stage, and the one connected to the first-stage is called the second-stage, and so on till the output system. In this thesis, the downstream classification is used for the force amplification leverage

- 10 -

mechanism. The upstream classification method is used for the displacement amplification leverage mechanism. A lever is used to trade a force with distance, or vice versa, to make a job function easier to do. The L/l value is called the lever ratio where L is the distance between the input and the pivot and l between the output and the pivot. More terminology for compliant leverage mechanisms will be presented later, such as amplification factor, mechanical advantage and geometry advantage, and mechanical efficiency.

2.4 Design Challenges

of

Microleverage

Mechanisms

and

Their

A microleverage mechanism differs in many aspects from that in the macro world. A pivot structure in the macro world can be formed by a pin-joint or bearing which permits free rotation and rigid support. Limited by current micro-fabrication technology, it is very difficult to achieve free rotation and rigid support in a microleverage mechanism. These elements are too complicated to be fabricated by surface micromachining. Some integrated fabrication processes, such as the SUMMIT process (The Sandia Ultra-planar Multi-level MEMS Technology,

http://www.mdl.sandia.gov/Micromachine) and MCNC (Markus and Koester, 1994, http://www.mcnc.org/), provide multiple polysilicon layers that can make pin-joints or bearings. However, at a pin-joint, the gap between the rotational and the fixed parts makes the mechanism behavior uncertain. Stiction force prevalent in the micro-world prevents a bearing pivot from functioning properly. Therefore, the most commonly used
- 11 -

pivot in MEMS is a flexure beam with one end anchored to the substrate. The geometry of the beam is designed to have a relatively small rotational spring constant allowing easy rotation. In a leverage mechanism, the lever arm needs to be kept rigid. The geometry and dimension of the flexure beams can be varied to have different spring constants. In most cases, even for a single-stage leverage mechanism, the magnification factor is a rough estimation. Many design issues and trade-offs are not well understood. For the compound leverage mechanisms, the geometry of each component is usually determined with a trial-and-error approach. The overall amplification factor has been obtained by testing and/or estimation. In many applications, the amplification factor needs to be accurately known. Taking a micro-tweezer for example, the force and displacement generated at the tip of the micro-tweezer need to be exactly known for picking up an object. Sensitivity of the resonant accelerometer, which is a precision instrument, needs to be precisely specified. In other cases where the input and output information is specified, a microleverage mechanism needs to be implanted to transfer the input system to the output system. It is imperative that an accurate expression of the amplification factor of the mechanism and a thorough understanding of the governing theory of the microleverage mechanism be developed. There are many design issues associated with microleverage mechanisms. The amplification factor depends on not only the geometry of the mechanism, but also the input and output system spring constants. In most cases, elastic deformation in the mechanism is small and the mechanism can be assumed to be operating in the linearelasticity regime. If the deflection is large enough to introduce non-linearity, the analysis

- 12 -

could become very complicated. The performance of the mechanism may also depend on the magnitude of the input force. Many factors influence the amplification factor, e.g., the pivot geometry, and how the leverage mechanism is connected to the input and output system. The stiffness of the output system plays an important role in a leverage mechanism. In a multiple-stage microleverage mechanism, the spring constant of the previous stage must be high enough to allow the next stage to have an amplification effect. It is similar to the design of an opamp in analog circuits where the input and output resistance of different stages should match. Designing a microleverage mechanism for a specific application can be a complex problem. In both single- and multiple-stage compound microleverage mechanism design, the challenging aspect is the conflicting stiffness requirement. The mechanism needs to be stiff enough to transfer the input force, yet soft enough to deflect. These issues will be addressed in detail in later Chapters.

2.5 FEM Simulation


For a single microlever, the magnification factor may be found from an analytical solution. As more and more stages are added, the analytical approach becomes more and more difficult and numerical methods are needed. While the analytical solution is based on many assumptions, the finite element method (FEM) requires fewer assumptions and is a more rigorous approach.

- 13 -

There are several software packages available for the FEM, such as ABAQUS (Hibbitt, 1994) and SUGAR (Clark, 1998). ABAQUS, together with PATRAN is a powerful tool for FEM simulation. It can be used to simulate fracture and PZT structures and to calculate resonant frequencies. It works for both linear and nonlinear elastic beam bending. Besides using PATRAN as a mesh generator, ABAQUS also has its own prior mesh generator and post analysis tool. However, ABAQUS requires substantial set up and it is difficult to interface the mesh generation for geometry variations. The results from ABAQUS are unnecessarily detailed for the analysis of microleverage mechanisms. SUGAR is a MEMS CAD tool developed at U.C. Berkeley (Clark, Zhou and Pister, 1998). It needs an input of a net-list file specifying the information of each connecting point as a node, which is similar to SPICE simulations of circuit. The netlist file is easy to change for geometry variations. SUGAR only simulates small deformation within linear elasticity which is the case for most microleverage mechanism designs. When a compliant mechanism undergoes large deflection and non-linearity is introduced, then a nonlinear FEM simulation such as ABAQUS must be used. Both ABAQUS and SUGAR are used for the single-stage microleverage mechanism simulation. The results are compared with each other and also with the analytical result. Figure 2.1 shows the mesh of the mechanism generated by ABAQUS. Figure 2.2 shows the deflected mechanism model. The ABAQUS input file for the simulation is attached in Appendix A. The geometry of the structure are as follows: horizontal beam length 210 m, width 20 m; middle beam length 24 m, width 2 m. The left beam functions as the pivot with length 6 m width 2 m. Distance between the

- 14 -

left and the middle beam is 10 m. With a input force 1.25e-9N, the average reaction force at the middle beam is 2.37e-8N from ABAQUS. From SUGAR the output force is 2.58e-8N and the amplification factor is 20. Results from ABAQUS and SUGAR are very close, with a discrepancy of 5%. Since in most of the application, the leverage mechanism deformations are in elastic regime, SUGAR is used in the microleverage mechanism simulation and is found to be quick and accurate. There are some discrepancies between the results from the 1st-order analytical model and the SUGAR simulation. With a second-order analytical model, the analytical result and the SUGAR simulation agree very well.

Fig. 2.1 Mesh generated by ABAQUS for the leverage mechanism.

Fig. 2.2 Leverage mechanism deflection simulated by ABAQUS.

- 15 -

CHAPTER 3 FLEXURE PIVOT DESIGN

An important aspect in the compliant microleverage mechanism is to replace the conventional pivot with a flexure hinge since a rotational pivot requires a more complex fabrication process. In a microleverage mechanism, the pivot is usually formed by a simple flexure beam with one end anchored to the substrate. In order to minimize stress concentrations at the corners where the pivot beam joins the lever arm, the joint usually has a rounded corner. Pivot design and optimization in microleverage mechanisms are presented in this Chapter.

3.1 Flexure Pivot Design


The classical paper on flexure pivot design is entitled How to design flexure hinges by Paros and Weisbord (1965). The paper states that while simple in shape and operation, flexures are mathematically complex. They extensively studied the design of the flexure hinges in several configurations and developed many approximate formulas. A general flexure hinge model is shown in Figure 3.1. The most common type of flexure hinge is basically a mechanical member which is compliant in bending about one axis but rigid about the cross axis. The term compliance is referred to as the reciprocal of the stiffness or the spring constant. In microleverage mechanisms, the flexure pivot is a very important component. The general flexure hinge model shown in Figure 3.1 can also minimize the stress concentrations. The flexure rigidity Kb (corresponding to a bending spring constant) and
- 16 -

the tensional rigidity Ks (corresponding to axial spring constant) are approximated by the following formulas (Her and Chang, 1994):

M 2 Eb t 5 / 2 = Kb = 9 r 1/ 2
Fx Eb

(3.1)

Ks =

(r / t )1/ 2 2.57

(3.2)

Where subscripts b and s denote bending and stretching, respectively. Both the analytical model and ABAQUS simulation of the hinges with different geometries are presented by Her and Chang (1994).

r Fx h
Fig. 3.1

b t

Fx

The generalized flexure hinge model.

Figure 3.2 and 3.3 show SEM of the pivots in a two-stage microleverage mechanism, which is fabricated with a Silicon-On-Insulator MEMS process, and is further described in Chapter 7. The first pivot is a generalized flexure hinge which serves as the first-stage pivot in the two-stage microleverage mechanism, Fig. 3.2. The other is a narrow beam which is the second-stage pivot in the two-stage microleverage mechanism, Fig. 3.3.

- 17 -

Fig. 3.2 A flexure pivot fabricated by SOI-MEMS technology.

Fig. 3.3 A flexure beam pivot in the second-stage microlever in the resonant Accelerometer fabricated by SOI-MEMS technology.

- 18 -

3.2 Different Models of Flexure Pivot


In microleverage mechanisms, most of the flexure pivots are formed by narrow beams. Figure 3.4 shows several prototypes with a flexure beam as the output: (a) the vertical model (model I), (b) the horizontal model (model II) and (c) the combined model (model III). The amplification factors of these three configurations are analyzed with SUGAR. A simple beam is used as the output and the output force is calculated as the axial tensile or compressive force. Table 3.1 lists the displacements and also the forces for these three models. The netlist files for geometry information are attached in Appendix B. The following parameters are used for the simulation: beam thickness, t, of 2 m; modulus of elasticity E = 1.65 x 1011 N/m2; lever arm length L = 200 m; distance between pivot and output system l = 10 m; input force, Fin, of 1.5 x 10-9 N; pivot beam and output beam width

wp = wo = 2 m; pivot beam and output beam length lp = lo = 60 m.


Compared with model II, model I has much greater amplification factor (See Table 3.1). Model III has an amplification factor in-between those of Model I and Model II. Figure 3.5 shows the comparison of the amplification factor of the each model as a function of the pivot beam length. Model I has the highest amplification factor compared to the other two and is not sensitive to variations of the pivot beam length. Model I will be used for all designs presented later. The vertical pivot is found to be the most effective for different pivot beam designs. The horizontal displacement and the rotation angle are slightly different when the pivot and output beams are at the same or different sides.

- 19 -

Table 3. 1 SUGAR Simulation Results of Different Pivot Models.

X Displacement Node 3 Y Displacement Rotational Angle X Displacement Node 4 Y Displacement Rotational Angle X Displacement Node 5 Y Displacement Rotational Angle X Displacement Node 2 Y Displacement Rotational Angle Amplification Factor

Model I 1.6556e-11 -2.6905e-13 5.5145e-07 1.6556e-11 2.8268e-12 5.523e-07 1.6556e-11 1.1556e-10 5.6934e-07 20.7

Model II 2.6395e-13 -7.878e-11 7.9263e-06 2.6615e-13 4.8811e-13 7.9273e-07 2.6615e-13 1.5882e-09 7.9443e-06 3.6

Model III 2.0073e-10 -1.3215e-10 1.3324e-05 2.0073e-10 1.0925e-12 1.3324e-05 2.0073e-10 2.6682e-09 1.3341e-05 1.1055e-13 -1.3167e-10 -1.607e-06 8.0

Table 3.2.

Comparison of SUGAR Results with the Model I Pivot and Output System on (a) the Same Side and (b) Opposite Sides of the Lever Arm Same Side 1.6556e-11 -2.6905e-13 5.5145e-07 1.6556e-11 2.8268e-12 5.523e-07 1.6556e-11 1.1556e-10 5.6934e-07 20.6 Different Sides 1.271e-14 -2.5857e-13 5.3048e-07 1.2858e-14 2.7221e-12 5.3134e-07 1.2858e-14 1.1126e-10 5.4838e-07 19.8

X Displacement Node 3 Y Displacement Rotational Angle X Displacement Node 4 Y displacement Rotational Angle X Displacement Node 5 Y Displacement Rotational Angle Amplification Factor

- 20 -

3.3 Pivot and Output System Configuration


Figure 3.6 is an interesting plot which shows how the amplification factor varies with the angle between the pivot beam and the lever arm. It is found that the amplification factor is maximum when the angle between the pivot beam and the lever beam is at 90 degrees. At 270 degrees, the amplification factor is slightly smaller than that at 90 degrees. When the pivot is at the same or opposite side of the output system, the mechanism is deflected differently. Sometimes, the micro-fabrication technology constraints will determine the leverage mechanism configuration, whether the pivot and the output system can be placed on the same side or different sides of the lever arm, e.g., SOI-MEMS technology only allows an outside anchor. Figure 3.7 shows the Model I pivot where the pivot and output system are on (a) the same side and (b) opposite sides of the lever arm. Table 3.2 lists the results from SUGAR simulation for a comparison between the two cases. The horizontal displacements are significantly different and the force direction changes. Also different are the rotation angles. Those differences cause slight amplification factor change for the single-stage leverage mechanism. Figure 3.8 shows the shape of the deflected pivot beam when the pivot and the output system are on different sides of the lever beam. Although the effect of the lever configuration on the amplification factor of a single-stage microleverage mechanism is insignificant, the effect becomes much more pronounced in a two-stage microleverage

- 21 -

mechanism. For a second-kind lever, the 2D microlever has a lower amplification factor than 2S, as explained in detail in later Chapters. The deflections of pivot and connection beams are simulated with SUGAR and plotted in Fig. 3.9. For illustrative purposes, the scale of the vertical axis (location on a beam) is 106 times that of the horizontal axis (deflection). When the connection beam and the pivot are on different sides of the lever arm, the shape of connection beam under loading is interesting, noting that they would have been bent in the same fashion if they had been on the same side of the lever arm. Because of the increased resistance to rotation when the pivot and output connection beam are on different sides of a lever arm, the amplification factor is reduced.

- 22 -

(a) Vertical Pivot


7 1 6 Output System 2 Pivot l 3 4 L 5

Input

(b) Horizontal Pivot


7 6 Output System

l 2 3 Pivot 4 L 5

Input

(C) Combined Pivot


7 6 11 1 2 Pivot l 4 L 5 Output System

Input

Fig. 3.4 Different Pivot Models.


- 23 -

25
Model I

20
2nd-Kind Microlever Lever Ratio 21:1

15
Model III

10
Model II

All dimensions in microns L= 200 l = 10 lo = 60 wo = wp = 2

0 0 10 20 30 40 50 60 70 80 90 100

Length of Pivot Beam, micron

Fig. 3.5

Comparison of the amplification factors of three Models as a function of the pivot beam length.
25
Model I 2nd-Kind Microlever Pivot Model I, II, and In-between Lever Ratio 21:1 All dimension in microns L= 200 l = 10 lo = 60 wo = wp = 2 Model I

20

Amplification Factor

15

10
Model II

0 90

110

130

150

170

190

210

230

250

270

Angle Between Pivot and Lever Arm

Fig. 3. 6

Effect of pivot beam with different angles on the amplification factor.

- 24 -

(a) Pivot at the same side as output

Output System

Pivot

Input

(b) Pivot at the different side as output

Output System

Input Pivot

Fig. 3.7

A second-kind leverage mechanism with the pivot and output system (connection beam) (a) on the same side and (b) different sides of the lever arm.

- 25 -

120 110

Location on the Pivot Beam, micron

100 90 80 70 60 50 40 30 20 10 0 0

Second-Kind lever with single beam as output All dimensions in microns L = 210 l = 10 lo = 100 lp = 120 wo = wp = 2

10

15

20

25

30

Horizontal Displacement, 1E-6 micron

Fig 3.8 The shape of a deflected pivot located on the opposite side of the lever arm.

Horizontal Displacement, 1E-6 micron


24 -6 -3 0 Node 6 20 Connection Beam

16

12

Node 7 Lever Arm

(exaggerated) 6 4 2 0 -6 -3 0 Node 8 Pivot Beam Node 9

Horizontal Displacement, 1E-6 micron

Fig 3.9 The deflection of pivot and connection beams of a 2D microlever.


- 26 -

CHAPTER 4 SINGLE-STAGE MICROLEVERAGE MECHANISMS

The single-stage microlever is the basic element in the microleverage mechanism. Analysis of the single-stage microlever provides insight and foundation for the design of both single- and multi-stage microleverage mechanisms. This chapter presents the analysis of single-stage microleverage mechanisms. A first-order analytical model and a second-order refined analytical model are built to derive the amplification factor, A, of the first-kind and second-kind microleverage mechanisms. The amplification coefficient, A*, is introduced to study the effect of the pivot geometry and output system on the amplification factor. Also presented are the strain energy analysis, mechanical efficiency, and mechanical and geometrical advantages of the mechanism. The design and optimization of the single-stage microleverage mechanism in a resonant output accelerometer is presented in detail. Other design issues, such as the effect of horizontal force and buckling are also addressed.

4.1 Basic Structure of A Single-Stage Microleverage Mechanism


The single-stage microleverage mechanism consists of four major parts: lever arm, pivot, the input system and output system, as schematically shown in Figure 4.1(a).

- 27 -

As a compliant mechanism consists of both rigid part and flexure members, the lever arm in a compliant microleverage mechanism is the rigid part while the pivot and the connections are the flexible elements. According to the different positions of the pivot, the input and output systems in relation to a lever arm, a microleverage mechanism can be classified as one of the three kinds. The first kind is defined by the pivot lying between the input and the output as in shown in Fig. 4.1(a). When the output lies in-between the pivot and the input, the microlever is then called a second kind as shown in Figure 4.1(b). When two lever arms of a second-kind lever join together by a hinge (i.e., pivot), it is then called a double second-kind lever. A third kind shown in Fig. 4.1(c) is defined as a microlever with the input lying between the pivot and the output. Joining the pivots of two third-kind microlevers makes a double third-kind lever, such as micro-tweezers. While first- and second-kind microlevers can amplify force, a third kind is typically used to magnify displacement. A first kind microlever can either amplify force or displacement, depending on the respective distances between (i) the pivot and the input and (ii) the pivot and output system. A first-kind microlever can also be used to change the force direction.

- 28 -

(a)First Kind
4. Output System

Anchor

3. Pivot l L 2. Lever Arm

1. Input System

(b) Second Kind


Output

l L

Input

(c) Third Kind


Output

l L

Input

Fig. 4. 1

Schematic of three kinds of micro-leverage mechanisms.

- 29 -

4.2 Amplification Factor of a Single-Stage Microleverage Mechanism


The leverage factor, l/L, is the maximum mechanical or geometrical advantage that a leverage mechanism can achieve. For force amplification, the leverage factor can also be called the amplification factor, A which is the output force v.s. the input force. The force amplification factor of a second-kind microlever is analyzed first. The analysis is similar for the other two kinds of microlever.

4.2.1 Second-kind single-stage microleverage mechanism


A second-kind microlever under loading is schematically shown in Fig. 4.2(a). When an input force is applied, the lever arm will be rotated by a small angle and displaced by a small distance, as shown in Fig. 4.2(a). The system under loading can be modeled as shown in Fig. 4.2(b). The flexural pivot is modeled as a kinematic joint, revolute pair, represented by a torsional spring and a vertical spring. For this analysis, five assumptions are made: (1) The elastic tension/compression or bending strain in the entire structure is within the regime of linear elasticity. (2) The lever arm is rigid and remains straight during loading. The total deflection can be represented by a vertical displacement (positive sign if the same direction as the input force, negative sign otherwise) and a rotation angle about the pivot (positive sign if counterclockwise).

- 30 -

(a)

Output Pivot Connection Beam to Output System

l +

Connection Beam to Input System

l L

(b)

Fig. 4.2

(a) A second-kind microlever before and after loading; (b) Model of the second-kind microlever under loading.

- 31 -

(3)

The un-anchored ends of the pivot beam and the connection beam to the output system maintains the 90 orientation with respect to the lever arm after loading. Therefore, the pivot beam and the connection beam will both be rotated by the same angle at the respective beam end.

(4)

The vertical output force does not cause any horizontal displacement or rotation of the output system.

(5)

The horizontal forces acting on the pivot beam and the output connection beam are negligible when a vertical input force is applied. In addition, the vertical displacement at output caused by the bending moment is assumed to be negligible.

Assumptions (3) to (5) imply that the rotation angle, , at the ends of the pivot beam and the output connection beam are solely caused by the respective bending moments. The vertical displacement at the output system is only caused by the vertical output force. Applying the force and moment (with respect to the joint between the pivot beam and the lever arm) equilibrium condition to the lever arm leads to the following equations:
F in = k vvo (l + ) + k vvp

(4.1) (4.2)

F in L = kvvo (l + ) l + k m o + k mp

where Fin is the input force, kvvo the vertical (first subscript v of the term kvvo) spring constant of the output (third subscript o) system under a vertical (second subscript v) force, kvvp the vertical spring constant of the pivot under a vertical force, k m o the

- 32 -

rotational spring constant of the output system when loaded with a bending moment,

k m p the rotational spring constant of the pivot when loaded with a bending moment, l
the distance between the pivot and the output and L the distance between the pivot and the input. The above equations were solved with Mathematica 4.0 and the Mathematica file used is attached Appendix C. Solving equations (4.1) and (4.2) for and by Mathematica, we have:

( k vvo + k vvp ) L k vvo l ( k vvo + k vvp )( km o + km p ) + k vvo k vvp l 2

Fin

(4.3)

( km o + km p) k vvo l ( L l ) ( k vvo + k vvp )( km o + km p) + k vvo k vvp l 2

Fin

(4.4)

Since the output vertical (i.e., axial) displacements caused by the internal bending moment and horizontal force are negligible, the output force is equal to the axial spring constant of the output system, kvvo, multiplied by the output axial displacement, + l. 1 F out = k vvo ( + l ) =

k vvp

( k mo + k m p ) + lL ( )
F in (4.5)

1 1 + + + l2 k vvo k vvp k m o k m p

The amplification factor of the microlever is:

- 33 -

1 A=

F out = 1 F in 1 + + + l2 k vvo k vvp k m o k m p

k vvp

( k mo + k m p ) + lL (

(4.6)

For an ideal lever, kvvp , k m p 0, k m o 0 and the amplification factor approaches the lever ratio:

Ao =

L l

(4.7)

4.2.2 First-kind microleverage mechanism


To derive the amplification factor of a first-kind microlever, the locations of the pivot and output system are exchanged with each other from those in a second-kind microleverage mechanism. For the amplification factor, we can just exchange kvvp with

kvvo, k m o with k m p in equations (4.1) and (4.2), and obtain the following:
1 A=

k vvp

( k mo + k m p ) lL ( )
(4.8)

1 1 + + + l2 k vvo k vvp k m o k m p

Using the same notation for L and l, the difference between the amplification factor of the first-kind and the second-kind microleverage mechanism is the sign of the lL term in the numerator. The first-kind microlever has a negative sign while the second-kind has a positive sign.

- 34 -

The first-kind and second-kind microleverage mechanisms are simulated with SUGAR for the purpose of comparison. It is found that the first kind had a slightly smaller amplification factor than the second kind for the same total length of the lever arm, L. This is because of the different lever ratios, i.e., L/l for the second kind and (Ll)/l for the first kind.

4.3 Amplification Coefficient, A*


The expression for the amplification factor in equation (4.6) does not give very much insight for the mechanism design. Another parameter, the amplification coefficient A*, which is more directly related to the design parameters, is defined as follows. The amplification factor of a microlever cannot be larger than that of an ideal lever, L/l. Taking into account that l<<L, the relative difference between the actual amplification factor and the ideal is:

1 l 1 + 1 k mo + k mp L k k vvo vvp A Ao = 1 Ao 1 + + + l2 k vvo k vvp k mo k mp

)
l <<1 L

1 1 + + k vvo k vvp k mo k mp (4.9) 1 1 + + + l2 k vvo k vvp k mo k mp

The numerator of the fraction expression on the right-hand side of equation (4.9) is now referred to as the microleverage mechanism amplification coefficient, A*. 1 1 + A*= + k vvo k vvp k m o k m p

(4.10)

- 35 -

The amplification coefficient is a product of (i) the sum of the rotational spring constants at the pivot and output system under a bending moment and (ii) the sum of the reciprocals of the axial (vertical) spring constants at the pivot and output system under an axial force. It can be seen from equations (4.9) and (4.10) that the amplification factor A decreases as the amplification coefficient A* increases and, more specifically,
A* 0, A = Ao , when A* << l 2 l2 1 1 A = Ao , when A* = l 2 = , 2 2 1, when A* >> l 2 A 0,

A* Ao - A Ao A* +l2

(4.11)

Therefore, for a single-stage microlever to have the maximum amplification factor, the amplification coefficient should be at a minimum, i.e., both kvvo and kvvp should be as large as possible and both k m o and k m p as small as possible. In other words, the pivot and output system need to be vertically (axially) stiff, but easy to rotate.

4.4

Effect of Output System and Pivot Spring Constants on the Amplification Factor
As seen from equations (4.9) and (4.10), among the axial and rotational spring

constants of the pivot and output system, kvvp , kvvo , k m p , and k m o , it is the larger rotational spring constant and smaller axial spring constant that has a more pronounced effect on the amplification coefficient. The vertical and rotational spring constants of the output system are typically known. Therefore, the pivot needs to have a large kv v p (kv v p>>kv v o) and a small k m p (k m p<<k m o). However, when k m p becomes less than

- 36 -

k m o, further decreases in k m p will not significantly increase the amplification factor


since the output rotational spring constant, k m o, becomes dominant.

The same reasoning holds for the vertical spring constant. Increasing the vertical spring constant of the pivot will increase the amplification factor. But when kvvp >> kvvo, the 1/kvvo term becomes dominant and further increases in kvvp will not significantly increase the amplification factor. Both kvvp and k m p are functions of the pivot

geometry. This explains why the length or width of the pivot has an optimum value for achieving maximum amplification factor as in the case of a resonant accelerometer presented in Chapter 7. In a case where the pivot is nearly perfect, i.e., large kv v p (kv v p>>kv v o) and small

k m p (k m p<<k m o), the amplification coefficient is approximately equal to k m o / kv v o,


according to equation (4.10). The amplification coefficient will increase and the amplification factor will decrease dramatically if the output system becomes more compliant, i.e., small kv v o . This conclusion is very important in the compound microleverage mechanism design, as will be illustrated in Chapter 5. In order for a downstream microlever stage to have a force-amplifying effect, the spring constant of the upstream stages, which are the output system for the downstream lever stage, cannot be too small. For a single-stage microleverage mechanism, the connection of the leverage mechanism to the output system also influences the force magnification factor. The

- 37 -

structure and the stiffness of the output system play an important role in the mechanism amplification factor. A leverage mechanism by itself may have a very appealing amplification factor, but it may drop dramatically when it is connected to an output system. For an ideal output system which is very stiff axially, but easy to rotate, in another word, experiencing little bending moment, i.e., kvvo and k m o 0, equation (4.10) gives

k m p wp 2 A* = 12 k vvp

(4.12)

The width of pivot, wp, is indeed a very critical dimension as far as the force amplification factor is concerned. For plotting graphs from the analytical equations, hypothetical values are given for the beam dimensions of the second-kind microlever. Substituting the following values in equation (4.6): Youngs modulus E = 1.65 x 1011 N/m2, beam thickness t = 2 m, the distance between the pivot and input system L = 210 m, the distance between pivot and output system l = 10 m, and computing the amplification factor as a function of outputsystem axial spring constant at a fixed output-system rotational spring constant of 2 x 107 Nm, the result is plotted in Fig. 4.3(a) and 4.3(b) for a series of pivot lengths and

widths, respectively. The effect of pivot width on the amplification factor is more pronounced than that of the pivot length. As seen from Fig. 4.3(a), the amplification factor is very low and

- 38 -

20

Amplification Factor, A

16

All dimension in microns L= 210 l = 10 lp = 6

12

lp = 20 lp = 30

8
lp = 40

4 Bending spring constant of the output system: 2E-7 0 1E+02 1E+03 1E+04 1E+05 1E+06 1E+07 1E+08

Output System Axial Spring Constant

Fig. 4.3

(a) Effect of output system axial spring constant on amplification factor for a series of pivot lengths.

20

Amplification Factor, A

16

All dimensions in microns L= 210 l = 10 lp = 6 wp = 1

12

wp = 2 wp = 3

wp = 4 wp = 5

4 Bending spring constant of the output system: 2E-7 0 1E+02 1E+03 1E+04 1E+05 1E+06 1E+07 1E+08

Output System Axial Spring Constant

Fig. 4.3

(b) Effect of output system axial spring constant on amplification factor for a series of pivot widths.

- 39 -

20

Amplification Factor, A

16

All dimensions in microns L= 210 l = 10 lp = 6 wp = 2

12

Output bending spring constant


2E-7 2E-6

8
2E-5 2E-4

0 1E+02

1E+03

1E+04

1E+05

1E+06

1E+07

1E+08

Output System Axial Spring Constant

Fig. 4.3

(c) Effect of output system axial spring constant on amplification factor for a series of rotational spring constants of the output system.
All dimension in microns L= 210 l = 10 lp = 6

20

Amplification Factor, A

16

12

Output axial spring constant


1E6

1E5 1E4

0 1E-08

1E-07

1E-06

1E-05

1E-04

1E-03

1E-02

Output System Bending Spring Constant

Fig. 4.4

Amplification factor as a function of the output system rotational spring constant for a series of output system axial spring constant.

- 40 -

increases slowly at low axial spring constants of the output system kv v o (kv v o << kv v p ), but increases sharply with increasing kv v o at intermediate values of kv v o and finally levels off at high value of kv v o (kv v o kv v p ). This observation can be explained from equation (4.6) and (4.10). The effect of the output system rotational spring constant, k m o, and axial spring constant, kv v o, on the amplification factor is shown in Fig. 4.3(c) and Fig. 4.4. At a fixed output system rotational spring constant, the amplification factor increases with an increasing axial spring constant of the output system, but levels off at a value which is dependent on the rotational spring constant of the output system. The higher the rotational spring constant of the output system, the lower the maximum achievable amplification factor. Figure 4.4 shows the dependence of amplification factor as a function of the output system bending spring constant for a series of output axial spring constant. It is found that the output system rotational spring constant can not be too big. The amplification factor drops dramatically as the output system rotational spring constant increases. Compliant mechanism performance can also be influenced by the input system compliance. This is because the motion of the mechanism is provided by the flexibility of all flexure hinges and linkages in the mechanism. However, for the analysis of microleverage mechanisms which function in the small linear elastic deformation regime, the effect of the input system spring constant and also the magnitude of the input force are neglected.

- 41 -

4.5 Second-Order Refined Analytical Model


The above analysis is based on five assumptions (Section 4.2.1). The fifth assumption, which ignores the horizontal forces, could cause some discrepancy between the analytical results and those calculated from FEM methods. A refined second-order analytical model is developed to take into account the horizontal displacement. The freebody diagram of a 2nd-kind microleverage mechanism is shown in Fig. 4.5. The rotation angle at the mobile end of the pivot beam, , is equal to the rotation angle of the rigid lever arm and can be separated into two components: m p (at pivot) caused by the bending moment and h p (at pivot) caused by the horizontal force. Similarly, the rotation angle at the joint between the output connection beam and the lever arm, , which is also equal to the rotation angle of the rigid lever arm, can also be separated into two components: the rotation angle m o (at output) caused by the bending moment and h o (at output) caused by the horizontal force. We have

= mo + ho

(4.13)

= mp + hp
Equilibrium of horizontal forces on the lever arm gives:

(4.14)

k ho ho = - k hp hp

(4.15)

where k h o and k h p are the rotational spring constants at output and pivot, respectively, when loaded with a horizontal force.

- 42 -

Since the lever arm remains rigid during loading and the vertical output force does not cause any horizontal displacement or rotation of the output system (the fourth assumption given before), we can have the following equation for the geometrical requirement of equal horizontal displacement at the two joints: (i) between the pivot and lever arm and (ii) between output connection beam and the lever arm:

k mp mp k hp hp k mo mo k ho ho + = + k hmo k hho k hmp k hhp

(4.16)

where k m o and k m p are the rotational spring constants at output and pivot, respectively, when loaded with a bending moment; k h m o and k h m p are the horizontal spring constants at output and pivot, respectively, when loaded with a bending moment;

k h h o and k h h p are the horizontal spring constants at output and pivot, respectively, when
loaded with a horizontal force. Replacing the total rotation angle in equation (4.2) with the rotation angle due to the bending moment, m o and m p , we have the following equation for the moment equilibrium with respect to the joint between the pivot beam and the lever arm:
F in L = k vvo l ( + l ) + k m o mo + k m pmp

(4.17)

Solving equations (4.13)-(4.16) for m o and m p , we have:

mo = f o mp = f p

(4.18) (4.19)

- 43 -

where fo and fp are functions of the eight spring constants in equations (4.13)-(4.16) and their expressions are cumbersomely long and not given here. However, for the special case of a resonant accelerometer, their expressions are much simpler and given later in equations. (4.39) and (4.40). Substituting equation (4.18) and (4.19) into equation (4.17), we can obtain:
F in L = k vvo l ( + l ) + k m o fo + k m p f p

(4.20)

Comparing equation (4.20) with (4.2), it is clear that k m o and k m p in equation (4.2) are now replaced by k m o fo and k m p fp in equation (4.20). The latter are now referred to as the apparent rotational spring constants, and fo and fp are the correction factors for the apparent rotational spring constants. In other words, the consideration of the horizontal forces at pivot and output is equivalent to replacing the actual rotational spring constants (with respect to a bending moment) with apparent ones. Furthermore, in this refined analysis, equation (4.1) from the first-order analysis still holds provided that the vertical displacement of the output system caused by the bending moment and horizontal force are negligible. The amplification factor and amplification coefficient can be obtained by following the same procedure in the first-order analysis. After solving and from equations (4.1) and (4.20), one can obtain the following equations:
1 A=

k vvp

( fo k mo + (

f p k mp ) + lL

1 1 f k + f p k mp + l 2 + k vvo k vvp o mo

(4.21)

- 44 -

(a)

Output Pivot

l +

l L

Input

(b)
Pivot

Output

Fhp Mp Fvp Fvp Mp Fhp

Fho Mo Fvo =Fout Fvo Fho Mo

Lever Arm Input Fin

Fig. 4.5 Free-body diagram of a second-kind microlever under loading.

- 45 -

1 1 f + A*= + f p k m p k vvo k vvp o k m o

(4.22)

It is noted that the above analysis is for a second-kind microlever. For a first-kind microleverage mechanism, For the single-stage microleverage mechanism design in a resonant accelerometer, results from both the first-order model and the refined secondorder model are plotted together with the SUGAR results, which are presented later in this Chapter. We will see that the refined model results are almost the same as the SUGAR results as shown in Section 4.9. This is a strong validation of both the SUGAR simulator and the analytical model.

4.6 Spring Constant Calculation


According to the beam bending theory (Timoshenko, 1974), the spring constants of a pivot beam fixed at one end with the other end subjected to axial (vertical) loading, lateral (horizontal) loading and bending moment, are called the axial spring constant

kvvp , lateral spring constant khhp, and bending spring constant kmp, respectively. These
spring constants are given as follows:
E wp t p lp 3E I p lp 3 EI p lp

kvp =

khp =

k p =

where I p =

t p wp 3
12

(4.23)

where E is the Youngs modulus, lp pivot beam length over which deflection occurs, tp

- 46 -

beam thickness, wp beam width, Ip moment of inertia about the neutral axis at midwidth(0.5wp is the distance from the neutral axis to the extreme fiber of the beam). For a general output system in MEMS devices, the spring constants kvvo, khho and

k m o , k h o can be calculated by using the unit load method. In a compound microlever,


the input spring constant of the upstream lever stage is the output spring constant of the downstream lever stage. The axial spring constant, of the entire lever at the input point can be calculated as:
2 k vo + k v p k o + k p + k vo k vp l F in = k in = 2 2 L + k o + k p + k vo (L l ) + k vp L

( (

)( )

(4.24)

For a second-kind single stage microlever, the relationship between the input and output force in an ideal case (the energies consumed at pivot and bending strain energy at output are not taken into account) obeys: Fout = Ao Fin. Ignoring the elastic bending energies consumed by the mechanism, the law of energy conservation gives Fin in = Fout out . We then have in = Ao out and kin = kout /Ao 2 . It is a special case of equation (4.24) noting that kout = kvvo. However, a compliant mechanism always consumes elastic bending energies, which is associated with the mechanical efficiency of the mechanism.

4.7 Strain Energy Analysis and Mechanical Efficiency


For a better understanding of the leverage mechanism and the influence of dimensional changes of pivot and the output system on the amplification factor, it is

- 47 -

necessary to analyze the strain energies consumed at the output system and the pivot when the input force is statically applied. In a compliant mechanism, the total strain energy is an integral of the distributed strain energy of the internal forces over the entire continuum. In general, the strain energy can be considered as the summation of the individual strain energies of a finite number of segments which idealize the continuum of the compliant mechanism. In the case of a microleverage mechanism, the internal strain energy consumed at the lever arm is negligible because it is relatively rigid and stays straight under loading. In the following first-order analysis, the strain energies consumed due to horizontal forces at the pivot and output are neglected. The work done by the input force is:

Uin =

Fin ( + L ) 2

(4.25)

The tensile/compression strain energy at the output system is: Uvo = Fout ( + l ) 2

(4.26)

The moment-bending strain energy at the output system is: M o 2

U o =

k m o 2
2

(4.27)

The tensile/compression strain energy at the pivot is:


Fvp 2

Uvp =

k vvp
2

(4.28)

The moment-bending strain energy at the pivot is:


- 48 -

U p =

Mp 2

k mp 2
2

(4.29)

The law of energy conservation gives:


U vo U o U vp U p + + + =1 Uin Uin Uin Uin

(4.30)

The total of strain energies passed to the output system and the energy consumed at the pivot should equal to the input energy. The major difference between rigid body mechanisms and compliant mechanisms is that energy is no longer conserved between the input and output port due to energy storage in the flexible parts of the latter. The ratio of the output to input energy is called the efficiency of the mechanism, e:

e = Fout out
F in in

(4.31)

Generally speaking, the higher the tensile/compression strain energy passes to the output system, the higher the ratio of the real amplification factor to the ideal one. It should also be noted that the ratio of tension/compression strain energy consumed at the output system to the input energy is not equivalent to the ratio of actual force amplification factor to the ideal one. Graphs of strain energies will be presented together with graphs of amplification factors for the case of a resonant accelerometer.

4.8 Mechanical Advantage and Geometrical Advantage

- 49 -

The mechanical advantage of a compliant mechanism is more complicated to quantify compared to rigid-link mechanisms whose mechanical advantage is well understood and readily evaluated. Due to component compliance, energy is consumed due to elastic deformation, and thus may not be assumed to be conserved between the input and output ports. Salamon (1998) presented the general relations for mechanical advantage of compliant mechanisms with single-input and single-output port using the energy method. In a compliant microleverage mechanism, the mechanical advantage, AM is the instantaneous ratio of the output force to the input force. It is of primary interest when the microleverage mechanism is designed for force amplification. Normally, mechanical advantage is maximized as the elastic deformation is minimized. AM = Fout /Fin (4.32)

The geometrical advantage, AG, is defined as the ratio of the output displacement to the input displacement. It is the amplification factor when the microleverage mechanism is used for displacement amplification. AG = out /in (4.33)

Apparently, if the elastic energy losses in a compliant mechanism are negligible, the product of AM and AG should equal to one. However, in compliant mechanisms the elastic energy loss due to elastic deformation is unavoidable. The product of AM and AG is equal to the mechanical efficiency of the mechanism which is always less than one. Sigmund (1997) discussed the mechanical advantage of a compliant mechanism with different
- 50 -

output configurations. The mechanical advantage strongly depends on the configuration and stiffness of the output system.

4.9 Single-Stage Microleverage Mechanism Design in a Resonant Accelerometer


Resonant inertial sensors are very attractive due to their frequency output and measurement precision. Resonant gyroscopes and accelerometers are designed and fabricated with integrated circuit technology. A microleverage mechanism is adopted in the resonant accelerometer to amplify the inertial force and increase the measurement sensitivity. Both anti-symmetrical and symmetrical structures have been designed. The following sections present the analysis and optimization of the symmetrical single-stage microleverage mechanism in the resonant accelerometer [Su and Yang, 200 (1)].

4.9.1 Resonant-output accelerometer (RXL)


A layout of the resonant-output micro-accelerometer is shown in Fig. 4.6 with four symmetrical single-stage second-kind microleverage mechanisms to amplify the inertial force of the proof-mass under acceleration or deceleration. It increases the sensitivity to more than 50 Hz/g from 3Hz/g of the first generation design, in which an asymmetrical leverage structure is used (Roessig, Pisano and Howe, 1995-1997). The proof-mass is the input system. The output system consists of two resonators (doubleended tuning forks) which vibrate at their natural frequency when actuated. Under acceleration or deceleration, the proof-mass generates a vertical inertial force. The

- 51 -

inertial force will be amplified by the microleverage mechanism before acting on the resonators. One of the resonators will be under tension and the other under compression. The axial loading changes the natural frequency of the two resonators, with tension force increasing it and compression force decreasing it. The frequency shifts of the two resonators are measured by an on-chip trans-resistance amplifier and off-chip electrical circuit to determine the acceleration. To analyze the leverage mechanism, Roessig used both FEM with ABAQUS to simulate the entire structure and also a much simplified analytical model to qualitatively estimate the amplification factor of the mechanism [Roessig, 1995]. The spring constant terms were omitted in the amplification factor. The following section provides an accurate and detailed analysis for the amplification factor and also the relationship between the amplification factor and the spring constant, more specifically the spring constant match theory. For the DETF resonator output system shown in Fig. 4.7(a), the axial spring constant can be simulated as that of a compound spring consisting of seven springs as shown in Fig. 4.7(b). Because of the symmetrical design of the DETF, its axial spring constant used in previous equations is actually the half of the axial spring constant of the entire tuning fork:
1 2k vvo = 1 + 1 1 1 1 1 1 + + + + k f2 + k f3 k f 4 k f5 k f6 + k f 7 k f2 + k f3 k f6 + k f 7

k f1

(4.34)

where beams 2 and 3 refer to tuning fork (T-F) beam, and 6 and 7 the connection (C-) beams between the tuning fork and the lever arm. The axial spring constants of these beams are:

- 52 -

k f 2 = k f3 =

E wf t f lf

kf 6 = kf7 =

E wc tc lc

(4.35)

where lf is the tuning fork beam length, tf tuning fork beam thickness, wf tuning fork beam width, lc joint beam length, tc joint beam thickness, wc joint beam width. Since the bending moments on the joint beams 6 and 7 in Fig. 4.7(a) cancel one another in beam 5 and do not get transmitted to beams 1-4, the moment-bending spring constant of the half tuning fork is:
E tc w c 3 12 lc

k mo =

(4.36)

For the second-order analysis in the case of the output system being a doubleended tuning fork, equations (4.15) and (4.16) become:
2 E Io lo 2 2 E Ip lp2

ho =

hp

(4.37)

lo 2

mo +

2 lo 3

ho =

lp 2

mp +

2 lp 3

hp

(4.38)

- 53 -

DETF Resonator (Output System)

T-F Connection Beam Lever Arm

Pivot and Anchor

Proofmass (Input System)

Fig.4.6 A layout of the resonant-output micro-accelerometer with a proof-mass, two resonators and four symmetrical single-stage second-kind microlevers.

- 54 -

Fig. 4. 7

(a) Schematic of DETF as the output system in the resonant accelerometer; (b) Simulated compound spring consisting of seven springs.

- 55 -

respectively. From equations (4.13), (4.14), (4.37) and (4.38), we can solve for m o and m p :
3(l o l p ) l o 2 wp3 l o3 wp3 + l p 3 wo 3

mo = f o

where f o = 1 +

(4.39)

mo = f p

where f p = 1 +

3(l p l o ) l p2 wo3 l p3 wo3 + l o 3 wp 3

(4.40)

Figure 4.8 shows the node information for the SUGAR simulation of the singlestage microleverage mechanism with the DETF output. An example of a netlist file for the SUGAR simulation is attached in Appendix D. The results are shown in the following sections.

4.9.2

Optimization of single-stage microleverage mechanism in resonant accelerometer


An analysis is given here for the amplification factor of a single-stage leverage

mechanism with prescribed dimensions of pivot and tuning fork beams and lever ratio. Beam thickness t, modulus of elasticity E, input force Fin, are the same as in Section 4.4. Substituting the values of the parameters mentioned above and the following values into the related equations: L = 210 m, l = 10 m, lf = 200 m, lj = 24 m, wj = 2 m, lp = 6 m, wp = 2 m, and computing the amplification factor and amplification coefficient as a function of tuning fork beam width wf , leads to the results as shown in Fig. 4.9(a). The amplification factors calculated from the SUGAR simulation are also plotted in Fig.

- 56 -

4.9(a) for comparison. For the given dimensions, the amplification factor increases continuously with increasing tuning fork width. On the other hand, the amplification coefficient decreases with increasing tuning fork width, as predicted by expression (4.10). The amplification factors calculated from SUGAR are close to, but always slightly lower than, those determined from the first-order analytical model. The secondorder refined analytical model is very close to the SUGAR result. The higher amplification factor predicted by the second order analytical model than SUGAR is mainly because the elastic deformation of the lever arm and other wide beams joining the tuning fork to the lever arm are ignored in the analytical equations, and the horizontal forces are not taken into account. The ratios of strain energies consumed at the tuning fork and at the pivot to the input energy are plotted in Fig. 4.9(b) as a function of tuning-fork beam width. As the tuning-fork beam width increases, the tensile/compressive strain energy at the tuning fork output (which is transferred from the input energy) increases gradually, leading to increases in the amplification factor.

- 57 -

1 2 3 DETF

4 5 12 DETF Connection Beam Lever Arm 13 17 14 8 16 10 6

15

9 7 11

Input

Fig. 4.8

The node information for SUGAR simulation of the single-stage microleverage mechanism in the resonant accelerometer

- 58 -

(a)
22
A, 1st-order analytical
A, 2nd-order analytical

1E-09
A, SUGAR

20

18

16

14
A*

All dimensions in microns L = 210 l = 10 lf = 200 lc = 24 wc = 2 lp = 6 wp =2

1E-10

1E-11

12

10 0 2 4 6 8

1E-12 10

T-F Beam Width, micron

(b)
1.0
Pivot Ten/Com Pivot Bending

Ratio of Strain Energy to Input Energy

0.8

T-F Bending

0.6

T-F Ten/Com

0.4

0.2

All dimensions in microns L= 210 l = 10 lf = 200 lc = 24 wc = 2 lp = 6

0.0

0.5

10

T-F Beam Width, micron

Fig. 4.9 Effect of the T-F beam width on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and the pivot to the input energy.

- 59 -

Amplification Coefficeint, A*

Amplification Factor, A

Similarly, the effect of the width of the T-F connection beam (connecting the T-F with the lever arm) on the amplification factor is shown in Fig. 4.10. For the given dimensions, the maximum amplification factor of 18.3 is found at a T-F connection beam width of ~ 1 m. However such width may be below the minimum beam width of 2 m from the lithography resolution. The ratios of strain energies consumed at the tuning fork and at the pivot to the input energy are plotted in Fig. 4.10(b) as a function of the T-F connection beam width. As the connection beam width increases beyond 1 m, the bending strain energy consumed at the T-F connection beam increases sharply, leading to lower amplification factors. It is seen from Figs. 4.9 and 4.10 that the width of T-F connection beam has a much more significant effect on amplification factor than the length of the tuning fork connection beam. This is because the former is undertaking the bending during loading while the latter is not, and the bending spring constant is proportional to the third power of beam width. The effect of pivot beam width on the amplification factor, as shown in Fig. 4.11(a), is very similar to that of the T-F connection beam. The results from the analytical equations and the SUGAR simulation match very well. The ratio of strain energies consumed at the tuning fork and at the pivot to the input energy is plotted in Fig. 4.11(b) as a function of pivot beam width. As the pivot beam width increases to a large value, the bending-moment energy consumed at the pivot increases sharply, leading to lower amplification factor.

- 60 -

(a)
20
A, 1st-order A,2nd-order analytical analytical

16

1E-09

A, SUGAR

12

1E-10

A*

1E-11

4 0 2 4 6 8

1E-12 10

T-F Connection Beam Width, micron

(b)
1.0
Pivot Bending

Ratio of Strain Energy to Input Energy

0.8

Pivot Ten/Com T-F Bending

0.6

T-F Ten/Com

0.4

0.2

All dimensions in microns L= 210 l = 10 lf = 200 lc = 24 wf = 2 lp = 6 wp = 2

0.0

0.5

10

T-F Connection Beam Width, micron

Fig. 4.10 Effect of the T-F connection beam width on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and the pivot to input energy.

- 61 -

Amplification Coefficeint, A*

Amplification Factor, A

All dimensions in microns L = 210 l = 10 lf = 200 lc = 24 wf = 2

1E-08

(a)
24 A, 1st-order analytical A, 2nd-order analytical A, SUGAR 16
All dimensions in microns L = 210 l = 10 lf = 200 lc = 24 wf = wc = 2

1E-08

20

1E-09

12

1E-10

8 A* 4

1E-11

0 0 1 2 3 4 5 6 7 8 9

1E-12 10

Pivot Beam Width, micron

(b)
1.0
Pivot Ten/Com Energy

Ratio of Strain Energy to Input Energy

0.8

T-F Bending Energy

Pivot Bending Energy

0.6
All dimensions in microns L= 210 l = 10 lf = 200 lc = 24 wf=wc = 2 lp = 6

0.4

T-F Ten/Com Energy

0.2

0.0

0.5

Pivot Beam Width, microns

Fig. 4.11 Effect of the pivot beam width on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and the pivot to the input energy.
- 62 -

Amplification Coefficient, A*
10

Amplification Factor, A

Figure 4.11(a) shows the amplification factor as a function of the tuning fork and connection beam width (tuning fork beam has the same width as that of connection beam) at various pivot beam widths, 0.5 5m. The general trend of the variation of amplification factor with increasing width of tuning-fork and connection beam is similar at different pivot beam widths. As the pivot beam width increases, the maximum amplification factor decreases and occurs at wider tuning fork and connection beams whose width roughly equals to the pivot beam width.

Pivot Width 20 1.0


1.5

Amplification Factor, A

16 2.0

12

2.5

3.0

8 3.5
4.0

4 4.5
5.0

All dimensions in microns L=210 l= 10 lf = 200 lc = 24 wf = wc lp = 6

0 0 1 2 3 4 5 6 7 8 9 10

T-F and Connection Beam Width, micron


Fig. 4.12 Amplification factor as a function of tuning-fork and connection beam width (tuning fork beam has the same width as that of the connection beam) at various pivot beam widths, 0.5-5 m.

- 63 -

With regard to the effect of the lengths of the tuning fork, T-F connection beam and pivot on the amplification factor, the results are shown in Figs. 4.13 and 4.14, 4.15, respectively. As shown in Fig. 4.13(a), a linear relationship exists between the amplification factor and tuning fork beam length because the tuning fork beams undertake no bending moments. By comparison, there is a sharp drop in amplification factor as joint beam and pivot beam length decrease to below 10 m, as shown in Figs. 4.14(a) and 4.15(a). This is due to the increased bending strain energy consumed at the TF joint beam and pivot beam as their lengths decrease as shown in Figs. 4.14(b) and 4.15(b). Compared with the effect of the beam width on the amplification factor, the beam length has much less effect. This is mainly because the bending spring constant is proportional to the third power of beam width, but the reciprocal of the beam length. Comparing the analytical and SUGAR results on the force amplification factor presented in Figs. 4.14(a) and 4.15(a), it is noted that the discrepancy between the two diminishes when the T-F connection beam length is equal to that of the pivot beam, i. e., 6 m in Fig. 4. 14(a) and 24 m in Fig. 4.15(a). The T-F connection beam can be regarded as a cantilever beam because its connection point with T-F does not experience any rotation or horizontal displacement. From the formula for the deflection and rotation of the angle of a cantilever beam with a bending moment applied at one end, it is known that x = l/2 where l is the beam length, x and are the horizontal displacement and rotation angle, respectively, of the pivot or the T-F connection beam caused by their respective bending moments. Therefore, the horizontal displacements of the two beams will be the same when (i) the rotation angle at the ends of these two beams caused by the

- 64 -

bending moments are the same as that of the lever arm, and (ii) they have the same length. In other words, in the case of a rigid lever arm, both geometrical requirements of (i) same rotation angle and (ii) same horizontal displacements of the pivot beam and T-F connection beam are satisfied simultaneously for the analytical solution when the two beams have the same length. Under this condition, no internal horizontal forces will have to be generated on those two beams. Otherwise, internal horizontal forces will be present in addition to the bending moments in order to cause the same rotation angle and the same horizontal displacement if those two beams are of different lengths. Ignoring these horizontal forces, the analytical solution will predict a higher force amplification factor than SUGAR. The effect of lever ratio, L/l, on the amplification factor is shown in Fig. 4.16 for a series of beam widths (for the sake of simplicity, all beams are assumed to have the same width). As expected, the amplification factor increases as the lever ratio increases. At large lever ratios, the amplification factor departs from that of an ideal lever and the difference increases as beam width increases. As lever ratio decreases, the amplification factor approaches that of an ideal lever, and the width of pivot and tuning-fork beam has less significant affect on the amplification factor. For a fixed lever ratio L/l of 21, the effect of l (the distance between pivot and tuning fork, or output arm length) on amplification factor is shown in Fig.4.17. As the output arm length l increases, the amplification factor approaches L/l which is the amplification factor of an ideal lever.

- 65 -

(a)
22
A, 1st-order analytical A,2nd-order analytical

1.0E-10

20

18

16
A* All dimensions in microns L = 210 l = 10 lc = 24 wf = wc = 2 lp = 6 wp = 2

1.0E-11

14

12

10 50

100

150

1.0E-12 200

T-F Beam Length, micron

(b)
1.0
Pivot Bending Energy

Ratio of Strain Energy to Input Energy

0.8

Pivot Ten/Comp Energy T-F Bending Energy

T-F Ten/Comp Energy

0.6

0.4

0.2

All dimensions in microns L= 210 l = 10 lc = 24 wf = wc = 2 lp = 6 wp = 2

0.0

50

75

100

125

150

175

200

T-F Beam Length, microns

Fig. 4.13

Effect of the T-F beam length on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and the pivot to the input energy.

- 66 -

Amplification Coefficient, A*

A, SUGAR

Amplification Factor, A

(a)
22
A, 1st-order analytical
A,2nd-order analytical A, SUGAR

1.0E-10

20

18

16
A*
All dimensions in microns L = 210 l = 10 lc = 24 wf = wc = 2 lp = 6 wp = 2

1.0E-11

14

12

10 50

100

150

1.0E-12 200

T-F Beam Length, micron

(b)
1.0
Pivot Ten/Comp Pivot Bending

Ratio of Strain Energy to Input Energy

0.8
T-F Bending

0.6
T-F Ten/Comp

0.4

0.2

All dimensions in microns L= 210 l = 10 lc =200 wf = wc = 2 lp = 6 wp = 2

0.0

10

20

30

40

50

60

T-F Connection Beam Length, microns

Fig. 4.14 Effect of the connection beam length on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and the pivot to the input energy.

- 67 -

Amplification Coefficient, A*

Amplification Factor, A

(a)
22.0
A, 1st-order analytical
A, 2nd-order analytical

1E-10

20.0

18.0

A, SUGAR

16.0
All dimensions in microns L = 210 l = 10 lf = 200 lc = 24 wf = wc = 2 wp = 2

1E-11

14.0

A*

12.0

10.0 0

10

20

30

40

50

1E-12 60

Pivot Beam Length, micron

(b)
1.00
Pivot Ten/Com

Ratio of Strain Energy to Input Energy

0.80
Pivot Bending
T-F Bending Energy

0.60
T-F Ten/Com Energy

0.40

0.20

All dimensions in microns L = 210 l = 10 lf = 200 lc = 24 wf = wc = 2 wp = 2

0.00 2 10

20

30

40

50

60

Pivot Beam Length, microns

Fig. 4.15 Effect of the pivot beam length on (a) the amplification factor and amplification coefficient; (b) the ratios of strain energies consumed at the tuning fork and the pivot to the input energy.

- 68 -

Amplification Coefficient, A*

Amplification Factor, A

60
All dimensions in microns L = 210 lf = 200 lc = 24 lp = 6 Ideal A

50

Amplification Factor, A

40

wf = 1 wc = 1 wp = 1

30

wf = 2 wc = 2 wp = 2 wf = 3 wc = 3 wp = 3

20

10 10

20

30

40

50

60

Lever Ratio, L/l

Fig. 4.16 Effect of lever ratio, L/l, on the amplification factor for a series of beam widths.

20

Amplification Factor, A

16

12

Fixed lever ratio of 21

All dimension in microns lf = 200 lc = 24 wf = wc lp = 6 wp=2

0 4 6 8 10 12 14 16 18 20

Distance between Pivot and T-F, micron

Fig 4.17

The effect of the distance between the pivot and the tuning fork, l, on the amplification factor for a fixed lever ratio L/l = 21.

- 69 -

20

18

Amplification Factor, A

16
All dimension in microns L = 210 l = 10 lf = 200 wf = 2 lc = 24 wc = 2 lp = 6 wp = 2

14

12

10 0 2 4 6 8 10 12 14 16 18 20

Lever Arm Width, micron

Fig. 4.18 Effect of the lever arm width on the amplification factor by SUGAR simulation.

Finally, the effect of lever arm width on amplification factor is shown in Fig. 4.18 where the data are taken from SUGAR simulations. It is found that, for a beam width of 2 m and other fixed dimensions, the amplification factor remains constant when the lever arm width is greater than 5 m. With a lever arm width greater than 20 micron, the lever arm can be considered rigid and have no influence on the amplification factor.

4.10 Other Design Issues of Microleverage Mechanisms


4.10.1 Beam column strength
For most micro-scale MEMS devices, the classical beam buckling theory still applies (Chiao and Lin, 2000). The critical load on a column fixed at one end and free at the other (Timoshenko and Goodier, 1970), is

- 70 -

P=

1 E I 2 , 4 l2

where I =

w 3t
12

(4.41)

which is one quarter of the strength of a column if both ends are hinged, 1/16 of the strength of a column if both ends are fixed. For a long and narrow beam with a width w of 2 m, thickness t of 2 m, length l of 200 m, and E = 1.65 x 1011 N/m2, the critical load is calculated to be 1.23 x 10-5 N. This determines the maximum allowable force in a microleverage mechanism before buckling. In the case of a resonant accelerometer with the single-stage leverage mechanism, for a total proof-mass volume of 500 x 500 x 2 m3 and silicon density of 2.3g/cm3, the accelerometer can sustain a maximum acceleration/deceleration of 6600 ms-2 as far as the T-F beam column strength is concerned. Flexure silicon beams in a microleverage mechanism usually fail by buckling before plastic deformation. Based on a yield strength of 7x109 N/m2, a beam width of 2 m and a beam length of 50 m, the maximum allowable axial load is 0.7N which are orders of magnitude higher than the applied load. For instance, in the resonant accelerometer, with a 1000g input acceleration and a leverage mechanism amplification factor of 200, the axial load on the DETF will be 2.3x10-4 N. At this load the beam will always be in the elastic regime, which is a basic assumption for SUGAR simulations. Similarly, worst case for a bending moment occurs with a 1000g input acceleration. This acceleration yields a force of 2.3x10-4N and given a 200 micron beam length a moment of 4.6x10-4Nm. The calculated allowable maximum bending moment in the beam is the yield strength multiplied by the section modulus, which is 2.33x10-7 Nm (0.233Nm),

- 71 -

orders of magnitude higher than the calculated bending moment. This shows that for most applications and typical loading conditions, the flexure beams are in the elastic region, therefore the SUGAR simulation is valid.

4.10.2

Effect of horizontal forces


If both vertical and horizontal loads act simultaneously on a beam, the resultant

stress or displacement at a point is equal to the superposition of the two separate effects. To balance an externally applied horizontal load Fh , the pivot and the output system would exert horizontal forces, Fh p and Fh o, respectively at their joint with the lever arm. The values of Fh p and Fh o can be calculated from the equation of static equilibrium and a geometrical requirement of equal horizontal displacements at beam ends of pivot and output system. It is clear that the external horizontal force does not generate vertical (axial) forces in the pivot or output system. Consequently, the effect of an external horizontal force is ignored in the analysis of a single-stage microlever.

- 72 -

CHAPTER 5 TWO-STAGE MICROLEVERAGE MECHANISMS

Single-stage microleverage mechanism can be very useful in MEMS applications to achieve significant mechanical and/or geometrical advantages. However, the amplification factor is strongly dependent on the leverage factor which is limited by the lever arm length allowable within a specified layout area. When an amplification factor greater than what a single-stage microlever can achieve is needed, a multistage compound microleverage mechanism can be considered. Compound levers have been used in the macro-world, such as for a weight machine (Beckwith, Buck and Marangoni). However, there are limited published designs of compound micro-levers, most being designed to amplify displacement. For example, Sniegowski and Smith (1995) at Sandia National Laboratory designed a leverage mechanism with a factor of 17.5 to convert a short-displacement, high-force electrostatic actuation to a long-displacement, medium-force output. Keller (1995-998) used a 2-stage microlever structure with an amplification factor of 27 in a micro-tweezer design to transfer a thermal expansion actuation to the motion of a micro-tweezer tip. However, the design theory for multiple-stage microlevers is still not well understood. By randomly stacking multiple levers together, the total force amplification factor could become

- 73 -

smaller and smaller. In other words, the added lever stage could have an amplification factor less than 1 with the upstream lever stage as its output system. The single-stage microlever is the basic element in a multistage microleverage mechanism. Building upon the analysis of a single-stage microlever in previous chapter, this Chapter presents the design theory of two-stage microleverage mechanism. The design theory for two-stage microleverage mechanisms is very important since it applies to the adjacent levers in multistage compound microleverage mechanisms with more than two stages. In most design cases, a two-stage microlever is more than adequate to achieve the desired amplification factor and it is indeed the most commonly used structure among compound microleverage mechanisms.

5.1 Two-Stage Micro-leverage Mechanism Structure


A two-stage compliant microleverage mechanism is formed by stacking two stages of levers together, with the upstream lever stage being the output system of the downstream lever stage, as schematically shown in Fig. 5.1. In a compound microleverage mechanism, the microlever stage connected to the output system through a connection beam is called the first-stage microlever. The microlever connected to the first is called the second stage, etc., until the last microlever stage connected to the input system. In order to avoid possible confusion between (i) the output system for the entire compound microlever and (ii) the upstream lever stage being the output for the adjacent downstream lever stage, the former is referred to as the external output system.

- 74 -

A classification scheme is described here for various two-stage microlevers according to the positions of input, pivot and output beams with respect to the lever arm of each lever stage. Using the same classification scheme established in Chapter 4 for single-stage microleverage mechanism, each lever stage in the two-stage leverage mechanism can be classified into different lever kinds depending on the relative location of the pivot, input and output beams on the lever arm. Note that the beam connecting the first stage microlever to the second stage is both the output beam of the second stage microlever and the input beam of the first stage. As described in Chapter 2, a leverage mechanism can be classified as one of three kinds. The first kind is the one with the pivot lying between input and output. When the output is between pivot and input, the microlever is called the second-kind. A third-kind is defined as a microlever with input positioned between pivot and output. A second-kind lever is for force amplification and third-kind for displacement amplification. A first-kind lever can amplify either force or displacement depending on the distance between the pivot to the input and the output. Depending on whether the output and pivot beams are at the same side or different sides of the lever arm, each kind of lever can be further classified into two subgroups. When the output and the pivot are on the same side of the lever arm, the lever is designated as S. If at different sides, the lever is represented as D. Based on the classification scheme mentioned above, the two-stage microlever shown in Fig.5.1 can be called a 1S-2S microlever. 1S before the dash sign represents that the first-stage microlever is a first-kind and S subgroup lever. 2S after the dash sign indicates the second stage microlever to be second kind and S subgroup.

- 75 -

Anchor Output System Fout , out Pivot l1 L1

First Lever Arm

F1, 1, K in, 1

Second Lever Arm Input Fin , 2, K in, 2 L2 l2

Fig 5.1

A schematic of a two-stage microleverage mechanism consisting of a first-kind lever as the first stage and a second-kind as the second stage.

Therefore, according to different kinds of lever (first- and second- kind) and different subgroups (S and D), a two-stage leverage mechanism for force amplification can have 16 (2x2x2x2) different configurations. The same is true for twostage leverage mechanisms used for displacement amplification (first- and third-kind, S and D subgroups). The amplification factors of the 16 different configurations are slightly different as presented in the following section.

5.2

Amplification Mechanisms

Factor

of

Two-Stage

Microleverage

- 76 -

The amplification factor of a two-stage microleverage mechanism is obtained by multiplying the amplification factors of the two single-stage microlevers. If the same assumptions are made for each stage of the two-stage microlever as in the case of a single-stage microlever discussed in Chapter 4, the total amplification factor for a twostage microleverage mechanism can be expressed as following: A = A1 A2 1 =

k vvp,1

( kmo,1 + km p,1 ) l1L1 ( )

k vp, 2

( kmo, 2 + km p, 2 ) l2 L2 ( )
(5.1)

1 1 + + + l12 k vvo,1 k vvp,1 km o,1 km p,1

1 1 + + + l2 2 k vvo, 2 k vvp, 2 km o, 2 km p, 2

where A1 and A2 are the amplification factor of the first- and second-stage microlever, respectively; kvvp, i and k m p, i (i = 1, 2) the vertical (axial) and the rotational spring constant of the pivot for the i microlever stage, respectively; kvvo, i the output vertical (axial) spring constant of the i
th th th

microlever stage, k m o, i the output rotational spring

constant of the i microlever stage, li the length between the pivot and output system for the i lever arm (positive sign for a second-kind microlever stage, and negative sign for a first-kind lever stage), Li the length between the pivot and input system of the i lever arm. The calculations of all the spring constants in equation (1) are given in the following. The axial and rotational spring constants of the pivot in each microlever stage are (i = 1, 2):
th th

kv vp , i =

E wp , i t p , i l p ,i

k m p, i =

E I p ,i l p ,i

where I p , i =

t p , i wp , i3
12

(5.2)

- 77 -

where E is the Youngs modulus of the microlever material, usually silicon; lp, i pivot beam length; tp, i beam thickness; wp, i beam width; Ip, i moment of inertia about the neutral axis at mid-width (0.5wp, i is the distance from the neutral axis to the extreme fiber of the beam). The axial/rotational spring constants at the output of the first-stage microlever can be calculated by treating the following two components as serially-connected springs: (i) the external output system and (ii) the beam connecting the external output system and the first-stage lever arm, i.e.,

kv vo, 1 =

1 1

k vvo

km o,1=

1 1

k vvc,1

k m o k m c,1

(5.3)

where kvvo and k m o are the vertical (axial) and rotational spring constants of the external output system for the two-stage microlever, respectively; kv v c, 1 and k m c, 1 the axial and rotational spring constants of beam C1 connecting the external output system and the first-stage lever arm, respectively. The other connection beam C2 is between the first and second microlever. The spring constants of the connection beams are:

kv vc,i =

E wc,i t c,i l c ,i

k mj ,i =

E I c,i l c,i

where I c,i =

t c,i wc,i3
12

(5.4)

where lc, i is the pivot beam length, tc, i beam thickness, wc, i beam width, Ic, i moment of inertia about the neutral axis at mid-width.

- 78 -

Similarly, the axial/rotational spring constants at the output of the second-stage microlever can be calculated by treating its output system as the first-stage microlever connected with the output system. It is noted that the rotation angle of the first-stage microlever caused by the vertical output force of the second-stage microlever is negligible as compared with the rotational angle of the second-stage lever arm. Then, an equation similar to equation (5.3) can be written for the second-stage microlever:

k v vo, 2 =

1 1

k vvI , 1 kvm c, 2

km o, 2 =

1 1

km I ,1 km c, 2

(5.5)

where kvvI, 1 and k m I, 1 are the input axial and rotational spring constants of the firststage microlever, and kv v I, 1 can be calculated as following,

2 kvvo, 1 + k vv p, 1 km o, 1 + k mp, 1 + k vvo,1 k vvp, 1 l1 F in = k vvI, 1 = 2 2 L11 + 1 km o, 1 + k m p, 1 + k vvo,1 (L1 l1 ) + k vvp, 1 L1

)(

(5.6)

k m I, 1 in equation (5.5) is calculated by treating the output system and the first-stage
pivot as two springs connected in parallel as far as sustaining a pure bending moment at the input point of the first-stage microlever is concerned, therefore,

k m I , 1 = k mo, 1 + k mp , 1

(5.7)

Substituting all the spring constants calculated from equations (5.2)-(5.7) into (5.1), the overall amplification factor can be obtained.

- 79 -

5.3 Compliance Relationship between Adjacent Levers


Designing compliant leverage mechanisms as mechanical amplifiers is analogous to designing operational amplifiers in analog circuits. If the force is analogous to the voltage and displacement to the current, then the spring constant corresponds to the resistance and the amplification factor corresponds to the gain. Similar to the impedance (resistance) matching requirement in the multistage op-amplifier to avoid degrading the overall gain, the compliance of different lever stages need to match in order for the entire mechanism to have a desired amplification effect. The key point in designing a two-stage microleverage mechanism is that the compliance need to match. As discussed in the Chapter 4, the amplification factor of a single stage microlever is dependent on the spring constant of the output system, a too soft output will cause the amplification factor to drop dramatically. The axial spring constant of the output has to be greater than a certain value in order for the mechanism to have an amplification effect. In a two-stage mechanism, the first-stage lever together with the external output will be the output of the second-stage lever. Therefore, not only the axial spring constant of the external output, but also the axial spring constant of the first-stage lever has to be greater than a critical value in order for the entire mechanism to have an amplification effect. Since the output axial spring constant of each microlever stage is a very critical parameter in determining the amplification factor of a two-stage microlever, the plots of amplification factor vs. output axial spring constant shown in Fig. 4.3 are re-plotted schematically in Fig. 5.2 for further discussion. This typically S-shaped plot can be

- 80 -

divided into three regions, I, II, III. Region I separates from Region II by Ai =1 (i = 1, 2 for first- and second-stage microlevers, respectively) when kvvo, i = k i. In Region I, the amplification factor is less than 1, which means the microleverage mechanism does not amplify force, but reduces it instead. k i can be calculated from equation (5.1) at Ai = 1,

ki ' = kv vo, i

A =1=
i

k mo, i + k m p, i
li Li li
2

kv vp

(k mo, i + k mp, i )

for first kind lever

(5.8a)

k i ' = kv vo, i

A i =1 =

k mo, i + k m p, i
li Li li 2

for second kind lever

(5.8b)

Ai

Ao, i

Amplification Factor of Ideal Microlever

Region III Ai @ Ao, i

Region II 1 Ai < Ao, i 1

Region I 1 Ai 

kvvo

Fig 5.2

Regime classification of the plot of the amplification factor as a function of the output system axial spring constants.

- 81 -

Region II is separated from Region III at an amplification factor close to the ideal one of Ao , i = Li / li when kvvo, i = k i. Assuming Ai = i Ao , i where i (referred to as amplification efficiency) is usually very close to the microlever energy efficiency ei defined later, k can be calculated from Eq. (1) at Ai = i Li / li ,

ki ' ' = k vvo, i

A i =i Li / l i

Li li (1 i )
2

Li i + Lii li kv vp,i

(5.9)

k mo, i + k m p, i

Region II is a workable region with the amplification factor ranging from 1 to

Ao , i . As the output axial spring constant increases from Region II to III, the
amplification factor approaches that of an ideal microlever. In Region III, further increase in kvvo, i does not cause any significant increase in the amplification factor. From equation (5.5) it is clear that the input axial spring constant of a microlever stage, kvvI, i , is approximately equal to the output axial spring constant of the adjoining downstream microlever stage, kvvo,i+1, because the axial spring constant of the connection beam, kvvc, i+1, is usually much greater than kvvI, i . Therefore, the axial spring constant at the input of an upstream microlever stage has a great influence on the amplification factor of the adjoining downstream microlever stage. When designing a two stage microleverage mechanism, the main aim is to make

kvvo, i greater than k for each microlever stage. It can be achieved by increasing kvvo,i or

- 82 -

making k smaller for a particular microlever stage. As can be seen from equation (5.8b), smaller k can be obtained by increasing lever arm length L, and decreasing k m o and

k m p. It should be noted that k m o is determined by the output system or the upstream


lever stage. After increasing the second stage pivot length or decreasing pivot width to decrease k m p to a certain degree, k m o becomes dominant and further increases in the pivot length or decreases in the pivot width will not significantly increase the amplification factor. Furthermore, for a two-stage microlever, the increased output axial spring constant of the second stage, kvvo, 2 , in Region III would result in insignificant increase in the amplification factor of the lever stage. At the same time, an increase in kvvo, 2 can only be caused by an increase in the stiffness of the first-stage, which would mean a pronounced increase in the amplification coefficient, A1 *, and a decrease in the amplification factor, A1 , of the upstream first-stage and the overall amplification factor of the two-stage microlever. While the second-stage microlever is optimized for its maximum amplification only, an optimized design of the first stage should be one which has a moderate amplification factor, A2 , and a relatively high input spring constant to allow the downstream stage to have a desirable amplification factor. Such compliance match between the two stages is the key in designing a two-stage microleverage mechanism for the maximum overall amplification factor.

- 83 -

As will be shown in the following analysis, the overall axial spring constant at the input becomes smaller and smaller when adding more and more upstream lever stages, with a reduction factor equaling to the square of the amplification factor divided by the leverage efficiency of the lever stage. This needs to be taken into account for the overall mechanism design. The mechanical efficiency of a microlever stage is defined as the ratio of output tension/compression strain energy to the input energy:

e1 = Fo o ,
F1 I ,1

e2 =

F1 I , 1 F 2 I ,2

(5.10)

where Fo is the output force (Fout), o vertical (axial) displacement at the external output system, F1 the axial force at the input of the first-stage microlever, F2 the axial force at the input of the second-stage microlever (which is also the axial input force of the entire two-stage microlever, Fin), I , 1and I , 2 are the vertical (axial) displacements at the input of first and second microlever stage, respectively. The ratio of output to input energy is always less than 1 because of tensile/compression and bending strain energy consumed at the pivot beams, connection beams, and the lever arm. From equation (5.10) and the definition of axial spring constant, we have:

kv vI , 1 =

F1

I ,1

(Fo o / e1 ) / F1 (Fo / F1) 2 / e1

F1

Fo / o

kv vo A12 / e1

(5.11a)

kv vI , 2 =

I ,2

F2

kv vI , 1 F1 / I ,1 F2 kv vo = = 2 = 2 F1 I ,1 / e2 / F 2 (F1 / F 2 ) / e2 A2 / e2 ( A1 A2 ) 2 / (e1e2 )

(5.11b)

- 84 -

It is clear from equations (5.11a) and (5.11b) that the overall axial spring constant at the input of a microlever stage is the output axial spring constant reduced by two factors: (i) the square of the amplification factor and (ii) the reciprocal of the energy efficiency of the microlever stage. In other words, the leverage mechanism becomes more compliant (i.e., the axial spring constant at the input becomes smaller) as more microlever stages are added. The output axial spring constant of the second-stage lever is the input axial spring constant of the first-stage lever. The input axial spring constant of the first-stage and second-stage can be calculated by equation (5.6). The input axial spring constant at microlever 2, kvvI, 2, is plotted in Fig. 5.3 as a function of the width of microlever 1 flexure beams (T-F, T-F connection beam, and pivot of the first-stage lever are all assumed to have the same width). Also plotted is the output axial spring constant, kvvo, divided by the square of (i) ideal total amplification factor, Ao (= Ao, 1 Ao, 2 ), and (ii) calculated total amplification factor, A (= A1 A2). It is seen that the geometrical average of kvvo /(Ao^2) and kvvo /(A^2) is almost the same as the input axial spring constant at microlever 2, kvvI, 2. In other words, the input axial spring constant of a compound microlever is approximately the output axial spring constant reduced by the ideal amplification factor and the actual amplification factor of the compound microlever. This is in agreement with equation (5.11) because Ao, 1 A1 / e1 and Ao, 2 A2 / e2.

- 85 -

100
All dimension in microns L= 200 l = 10 lf1 = 100 lc1 = lp1 = 6 lc2 = lp2 = 60 Kvvo/(A^2)

80

Axial Spring Constants, K

60

40
Geo. Ave.

20
KvvI, 2 Kvvo/(Ao^2)

0 1 2 3 4 5 6 7 8 9 10 11 12

Beam Width of Pivot & T-F, micron

Fig. 5. 3

Input axial spring constant of the second stage and the calculated values of

kvvo /(Ao^2) and kvvo /(A^2) and the geometrical average of the two.

k vvo
2 Ao

k vvI , 2

k vvo A2

(5.12)

k in,1

k vvo k vvo
2 A 2 Ao

k vvo Ao A

(5.13)

The compliance contradiction, which is common in compliant mechanism design, can also be seen here. In order for the second stage microlever to have bigger amplification factor, the spring constant of the first stage microlever must be high. But if the first stage lever has a high amplification factor, then the input spring constant would be low. The design of two-stage microleverage mechanism is a compromise of each stage

- 86 -

amplification factor for achieving an optimum overall amplification factor. The other compliance contradiction lies in that the mechanism needs to be compliant enough to allow the maximum input energy at the input port, while the mechanism has to be axially stiff enough for each stage to have the maximum amplification effect.

5.4

Different Mechanisms

Configurations

of

Two-stage

Leverage

As classified in section 5.2, there are sixteen different configurations of two-stage leverage mechanisms for force amplification and sixteen different configurations for displacement amplification. They are distinguished by different kinds of lever combinations (the first- and second-kind for force amplification, the first- and third-kind for displacement amplification) and different sides of the pivot and output locations (S for same and D for different). For the first- and second- kind of levers there is a sign difference (+ or -) in the analytical expressions of the amplification factor. However, for the subgroup of pivot and output being on same and different sides of the lever arm, the first-order analytical amplification factor can not be distinguished. That is because the analysis is based on the assumptions given before without taking into account the strain energy consumed on the lever arm and the internal horizontal forces acting on the pivot and connection beams. Table 1 lists all sixteen configurations for a two-stage leverage mechanism for force amplification and their deflections after loading. Also listed are the rotation angles of the two lever arms and the amplification factor of the mechanism. The starting (i.e.,

- 87 -

before loading) positions of the lever arm and flexure beams are shaded and the afterloading positions shown with solid lines (exaggerated for illustration purposes). The third and fourth columns of Table 1 list the rotation angles of lever arms and the forceamplification factors of first- and second-stage, respectively. The pivot and connection beam are sometimes deflected into a twisted curve when the output and the pivot are on different sides of the lever arm, contributing to the lower amplification factor than the S group. The dimensions used for SUGAR simulation are as follows, with a double-ended tuning fork as the output: lever arm distance between input and pivot L1 = L2 = 200 m, between pivot and tuning fork l1 = l2 = 10 m (negative sign for first-kind microlevers), width of lever 1 and lever 2 pivot beam, and connection beam between two microlever stages wp2 = wp2 = wc2 = 2 m, length of the tuning fork beam lf = 100 m, microlever 1 pivot lp1 = 6 and T-F connection beams lc1 = 6 m, length of microlever 2 pivot, and connection beam between two microlever stages

lp2 = lc2 = 60 m.
Figure 5.4 shows the comparison of the amplification factor of two-stage leverage mechanisms of different configurations. The two-stage levers with the pivot and output on the same side of the lever arm (e.g., 2S-1S, 2S-2S, 1S-1S and 1S-2S configurations) have much higher amplification factor than those with the pivot and output

- 88 -

Table 5.1. Beam Deflections and Amplification Factor of Two-stage Microlevers NO. 1 CODE TWO-STAGE LEVER
CONFIGURATION

ROTATION,  1 = 4.21 x 10- 7 2 = - 8.36 x 10- 6

LEVER ARM

AMPLIFICATION FACTOR
EACH STAGE A1 = 5.9 A2 = 3.8 TOTAL A = 22

1D-1D

1D-2D

1 = 4.45 x 10- 7 2 = - 8.67 x 10- 6

A1 = 6.6 A2 = 3.5

A = 23

1D-1S

1 = 1.16 x 10- 6 2 = -2.24 x 10- 5

A1 = 6.4 A2 = 9.4

A = 60

1D-2S

1 = 1.09 x 10- 6 2 = - 2.12 x 10- 5

A1 = 6.1 A2 = 9.5

A = 57

1S-1D

1 = - 4.47 x 10- 7 2 = - 8.99 x 10- 6

A1 = 11.0 A2 = 2.5

A = 27

1S-2D

1 = 4.78 x 10- 7 2 = - 9.44 x 10- 6

A1 = 16.7 A2 = 2.1

A = 34

1S-1S

1 = -1.43 x 10- 6 2 = -2.82 x 10- 5

A1 = 12.9 A2 = 6.7

A = 86

1S-2S

1 = 1.33 x 10- 6 2 = -2.64 x 10- 5

A1 = 11.7 A2 = 6.9

A = 80

- 89 -

Table 5.1. Beam Deflections and Amplification Factor of Two-stage Microlevers (continued) CODE 9 TWO-STAGE LEVER
CONFIGURATION

LEVER ARM

AMPLIFICATION FACTOR
EACH STAGE A1 = 6.2 A2 = 3.9 TOTAL A = 24

ROTATION, 

2D-1D

1 = -4.20x10-7 2 = -8.26x10-6

10

2D-2D

1 = 4.44x10-7 2 = -8.60x10-6

A1 = 6.8 A2 = 3.7

A = 25

11

2D-1S

1 = -1.13x10-6 2 = -2.19x10-5

A1 = 6.6 A2 = 9.7

A = 64

12

2D-2S

1 = 1.07x10-6 2 = -2.07x10-5

A1 = 6.3 A2 = 9.7

A = 61

13

2S-1D

1 = -4.51x10-7 2 = -9.03x10-6

A1 = 11.2 A2 = 2.4

A = 26

14

2S-2D

1 = 4.82x10-7 2 = -9.51x10-6

A1 = 14.1 A2 = 2.0

A = 28

15

2S-1S

1 = -1.46x10-6 2 = -2.87x10-5

A 1 = 13.2 A2 = 6.4

A = 84

16

2S-2S

1 = 1.36 x 10-6 2 = -2.68 x 10-5

A1 = 11.8 A2 = 6.7

A = 79

- 90 -

80

A1

A2

Total A

Amplification Factor

60

40

20

0 2D-1D 2D-2D 2D-1S 2D-2S 2S-1D 2S-2D 2S-1S 2S-2S

Types of Two-stage Microlever

80

A1

A2

Total A

Amplification Factor

60

40

20

0 1D-1D 1D-2D 1D-1S 1D-2S 1S-1D 1S-2D 1S-1S 1S-2S

Types of Two-stage Microlever

Fig. 5.4

Comparison of the amplification factors of two-stage microlevers of different configurations.

- 91 -

on the different sides of the lever arm. The increased resistance of a microlever to rotational and axial displacement when the output and pivot are on the different sides of a lever arm leads to lower force-amplification factor. Among the sixteen configurations, the second-stage pivots in 1D-1S, 1S-1S, 2D-1S and 2S-1S are inside the enclosure formed by the connection beam, first-stage and second-stage lever arm. Therefore, the pivot can not be longer than the connection beam. Such a constraint limits the amplification factor and so those four configurations are not commonly used. In summary, the optimum design of two-stage microlever should be of 1S/2S1S/2S types of combinations, i.e., the pivot and connection beam of both the first- and second-stage microlever are on the same side of the lever arm. In addition, stiffness (axial spring constant) of the first-stage microlever (connected to output system) should be significantly greater than that of the second-stage. A wider gap between the first- and second-stage microlever would allow the use of very long pivot and connection beams for the second-stage microlever, which can reduce its axial spring constant and increase the force-amplification factor.

5.5

Two-Stage

Microleverage

Mechanism

Design

and

Optimization in the Resonant Accelerometer


The previous chapter presents the design and optimization of the single-stage microleverage mechanism in the resonant accelerometer. This section will present the

- 92 -

design and optimization of two-stage microleverage mechanism in the resonant accelerometer [Su and Yang, 2000 (2)]. Among the sixteen configurations of the two-stage leverage mechanism for force amplification, four of them (1D-1S, 1S-1S, 2D-1S and 2S-1S) have inside second-stage pivots which limit the amplification factor. Among the other twelve configurations of the two-stage leverage mechanism, six configurations are feasible for the resonant accelerometer application as shown schematically in Figs. 5.5 (a) to (f). The area between the two lever stages is used as a portion of the proof-mass. Three configurations (1S-2S, 2S-2D and 2S-2S) have inside anchors. For the optimization of the two-stage microleverage mechanism, the amplification factors of a 1S-1S type are calculated with prescribed dimensions of pivot, tuning fork beams, connection beams, and lever ratio. For a symmetric two-stage lever design and a constant beam thickness, t, of 50 m, substituting the following values into equations (5.2) (5.7) for calculating spring constants and finally into equations (5.1) for calculating the amplification factor: modulus of elasticity E = 1.65 x 1011 N/m2, input force, Fin, = 1.5 x 10-9 N, lever arm distance between input and pivot L1 = L2 = 210 m, lever arm distance between pivot and tuning fork l1 = l2 = 10 m, tuning fork beam length lf = 100 m,

- 93 -

1 2 3

4 12 16 15 14 5 6 10 9 13 7 8 11

17

18 19

20

23

24 25

Proof-mass 21 22 26 27

Fig. 5.5 (a) A 1S-1D type twostage microleverage mechanism.

1 2 3

4 12 16 15 17 14 13 7 8 5 6 10 9 11

19 18

20

23

25 24

Proof-mass 21 22 27 26

Fig. 5.5 (b) A 1S-2D type two-stage microleverage mechanism.

- 94 -

1 2 3

4 12 22 21 17 16 15 14 5 6 10 9 13 7 8 11 27 26

19 18

20

23

25 24

Proof-mass

Fig. 5.5 (c) A 1S-2S type two-stage microleverage mechanism.

1 2 3

4 12 15 17 14 13 8 7 5 6 9 11

16 10

18 19

20

23

24 25

Proof-mass 21 22 26 27

Fig 5.5 (d) A 2S-1D type two-stage microleverage mechanism


- 95 -

1 2 3

4 12 15 17 13 14 5 6 9 7 8 11

16 10

19 18

20

23

25 24

Proof-mass 21 22 27 26

Fig 5.5 (e) A 2S-2D type two-stage microleverage mechanism.

1 2 3

4 12 22 21 17 15 13 14 8 7 5 6 27 9 11 26

16 10

19 18

20

23

25 24

Proof-mass

Fig. 5.5 (f) A 2S-2S type two-stage microleverage mechanism.

- 96 -

width of lever 1 pivot, T-F beams, and T-F connection beams wp1 = wf = wc1 = 4 m, or as variables, width of lever 2 pivot, and connection beam between lever 1 and 2 wp2 = wc2 = 2 m, or as variables, length of lever 1 pivot and T-F connection beams lp1 = 6 and lc1 = 24 m, or as variables, length of lever 2 pivot, and connection beam between lever 1 and 2 lp2 = lc2 = 60 m, or as variables, the amplification factor of individual lever, A1 and A2, and total amplification factor, A, can be calculated and the results are shown in Figs. 5.6 5.11. The same dimensions are used in SUGAR simulation for comparison with the analytical results in Fig. 5.9. The amplification factors of the individual levers, A1 and A2, and the total amplification factor, A, as a function of beam width of tuning fork, connection beam and pivot of lever 1 (the width of all beams are assumed to be same) are shown in Fig. 5.6(a) and 5.6(b). The width of the first-stage lever pivot, T-F beam and connection beam are chosen to be 4 m to increase the stiffness of the first stage lever to ensure that lever 2 will have a good amplification factor. For the same reason, the lengths of lever 1 pivot and T-F connection beams are 6 m, which is relatively short compared to those of the second stage lever. As the beam width in lever 1 increases, the amplification factor of lever 1, A1, decreases while the amplification factor of lever 2, A2, increases. An increase

- 97 -

in the amplification factor of lever 1 results from an increase in the axial output spring constant of lever 1 when the tuning fork, connection beam and pivot beam width in lever 1 increases, as shown in Fig. 5.6(b). A maximum amplification factor is obtained at an intermediate beam width of ~4 m. The separate effects of the width of three flexure beams in lever 1: (i) tuning fork beam, (ii) connection beam, and (iii) pivot beam on amplification factors of both lever 1 and 2 and the total amplification factor are shown in Fig. 5.7(a)-(c), respectively. The continuous increase in the amplification factors of both lever 1 and 2 and the total amplification factor with increasing width of the lever 1 T-F beam, shown in Fig. 5.7(a), is expected because (i) the T-F beam consumes no bending energy; (ii) wider T-F beam means greater vertical output spring constant for lever 1; (iii) wider T-F beam also results in a greater vertical output spring constant for lever 2. However, the greater width of the tuning fork may adversely affect the measurement sensitivity. The amplification factor of lever 1 and the total amplification factor are almost constant for a T-F connection beam width of less than 4 m, but then decrease rapidly with wider T-F connection beams, as shown by Fig. 5.7(b). Similar effects on the amplification factor are seen in Fig. 5.7(c) for pivot beam width. The decrease in amplification factor of lever 1 at wider widths of the T-F connection beam or the pivot beam is due to increased moment bending energy consumed at these two beams as their width increases. Although an increase in the width of the T-F connection beam and the pivot beam of lever 1 causes an increase in the vertical output spring constant of lever 2 and, consequently, an increase in the amplification factor of lever 2, the decrease in the

- 98 -

amplification factor of lever 1 overrides the increase in the amplification factor of lever 2. This leads to a decrease in the total amplification factor of the compound lever as the width of the T-F connection beam and the pivot beam of lever 1 increases. The effects of the lengths of T-F, connection beam and lever 1 pivot, on the amplification factor are shown in Fig. 5.8 (a), (b) and (c) respectively. Comparing Fig. 5.8 with Fig. 5.7, it is seen that the effect of flexure beam length on amplification is less significant than the effect of beam width. As the length of T-F beam increases, the amplification factors of both lever 1 and 2 decrease because of decreasing vertical output spring constants, kvvo, 1 and kvvo, 2, of both levers. Increasing the length of the T-F connection beam causes less drop in the vertical output spring constant than increasing the width. The amplification factor decreases only slightly as the T-F connection beam length increases after an initial increase to a maximum at relatively short beam length. Comparing Fig. 5.8(b) with 5.8(c), it is seen that the length of the lever 1 pivot beam has almost the same effect on amplification factor as the length of the T-F connection beam. A SUGAR simulation of the results of two different types of two-stage microlevers, 1S-2S and 1S-2D, are presented in Fig. 5.9 for comparison with the 1st-order analytical results as shown in Fig. 5.7(c). The amplification factors calculated from the first-order analytical model are slightly higher in the case of the 1S-2S lever and much higher in the case of 1S-2D lever than the SUGAR results. The first-order analytical model gives reasonable predictions of the amplification factor when the pivot and output system are on the same side of the lever arm in both the first- and second-stage levers. When the pivot and the output system are on different sides, different horizontal force

- 99 -

generates different amplification factor. The 1st-order analytical model can not show the difference. However, for a two-stage leverage mechanism, the 2nd-order analytical model is too complicated to pursue. To distinguish the subgroup of S (pivot and output are at the same side) and D (pivot and output are at different side) in the two-stage leverage mechanism, the SUGAR simulation is the main approach here since all of the beam deformation is in elastic regime. The effect of the width of lever 2 flexure beams on amplification factor is shown in Figs. 5.10(a) and (b) for the lever 2 connection beam and pivot beam, respectively. The width of the lever 2 connection beam has the same effect on the amplification factor as the pivot beam. A comparison between Fig. 5.11 and Fig. 5.10 shows that the width of the lever 2 flexure beams has more significant effect on total amplification factor than that of the lever 1 flexure beams. This is a result of sharp decreases in the amplification factor of lever 2 at increasing width of lever 2 flexure beams. Figure 5.11 shows the effect of the length of the lever 2 connection beam on the amplification factors, which is opposite to the effect of the width of lever 2 flexure beams shown in Fig. 5.10. While increasing the length of the lever 2 connection beam has no effect on the amplification factor of lever 1, it has a pronounced effect on the amplification factor of lever 2 and the total amplification factor. The effect of the length of lever 2 pivot beam on the amplification factors has been found to be the same as that of the length of the lever 2 connection beam. All the results presented in Figs. 5.6 5.11 agree well with the compliance analysis given in Section 5.3. In all, the stiffness (axial spring constant) of the first-stage

- 100 -

microlever (connected to output system) should be significantly greater than that of the second-stage. A wider gap between the first- and second-stage microlever would allow the use of very long pivot and connection beams for the second-stage microlever, which can reduce its axial spring constant and increase the force-amplification factor.

- 101 -

180
First-kind

160 140

Amplification Factor

120 100 80 60 40 20 0 1 2 3 4 5 6 7 8 9

All dimension in microns L = 200 l = 10 lf1 = 100 lc1 = lp1 = 6 lc2 = lp2 = 60 Total w2=1.5 w2=2.0 w2=2.5 A2 w2=1.5 w2=2.0 w2=2.5 A1

10

11

12

Beam Width of Pivot & T-F, micron

Fig. 5.6 (a) Amplification factors, A1, A2 and A, as a function of the width of lever 1 TF, T-F connection beam, and pivot beam (SUGAR).
40
First-kind Levers

1E+04
All dimension in microns L = 210 l = 10 lf1 = 100 lc1 = lp1 = 6 lc2 = lp2 = 60

Amplification Factor of 2nd Stage, A2

35 30 25 20 15 10 5 0 1

A2 w2=1.5 1E+03 w2=2.0 w2=2.5

1E+02 10 11 12

Lever 1 Beam Width, micron

Fig. 5.6

(b) First-stage lever spring constant k and second-stage lever amplification factor as a function of lever 1 flexure beam width (SUGAR).

- 102 -

Lever 2 Output Axial Spring Constant

180
Total A

160 140

Amplification Factor

120 100 80 60 40 20 0 1 2 3 4 5 6

First-kind Levers All dimension in microns L = 210 l = 10 lf1 = 100 lc1 = lp1 = 6 wc1 = wp1 = 4 lc2 = lp2 = 60 wc2= wp2 = 2

A1 A2

10

11

12

Width of T-F Beam, micron

Fig. 5.7 (a) Amplification factors, A1, A2 and A, as a function of the lever 1 T-F width.
180
First-kind Levers

160 140

Amplification Factor

120 100 80 60

All dimension in microns L= 210 l = 10 lf1 = 100 lc1 = lp1 = 6 wf = wp1 =4 lc2 = lp2 = 60 wc2 = wp2 = 2

Total A

40 20 0 1 2 3 4 5 6 7 8 9 10
A2 A1

11

12

Width of T-F Connection Beam , micron

Fig. 5.7 (b) Amplification factors, A1, A2 and A, as a function of the lever 1 T-F connection beam width (SUGAR).

- 103 -

180
First-kind Levers

160 140
All dimension in microns L= 210 l = 10 lf1 = 100 lc1 = lp1 = 6 wf = wc1 = 4 lc2 = lp2 = 60 wc2 = wp2 = 2

Amplification Factor

120 100 80 60

Total A

40 20 0 1 2 3 4 5 6 7 8 9 10
A2 A1

11

12

Width of Pivot Beam, micron

Fig. 5.7 (c) Amplification factors, A1, A2 and A, as a function of the lever 1 pivot width.
180
First-kind Levers

160 140
All dimension in microns L= 210 l = 10 lf1 = 100 lc1 = lp1 = 6 wf = 4 lc2 = lp2 = 60 wc2 = wp2 = 2

Amplification Factor

120 100 80 60 40 20 0 1 2 3 4 5 6 7

Total A

A2 A1

10

11

12

Pivot & Connectio Beam Width, micron

Fig. 5.7

(d) Amplification factors, A1, A2 and A, as a function of width of the lever 1 pivot and connection beam (SUGAR).

- 104 -

160 First-kind Levers 140 120 100 80 60 40 20 0 50 Kvo,2 A1 A2 Total A


All dimension in microns L = 210 l = 10 wf= wc1 = wp1 = 4 lp1 = lc1 =6 wc2 = wp2 = 2 lc2 = lp2 = 60

1E+10 1E+09 1E+08 1E+07 1E+06 1E+05 1E+04 1E+03 1E+02 250

Kvo,1

100

150

200

Lever 1 T-F Beam Length, micron

Fig. 5.8 (a) Amplification factors, A1, A2 and A, as a function of the T-F beam length.
160 First-kind Levers 140 120 100 80 Kvo,1 60 40 20 0 0 10 20 30 40 50 60 70 80 Kvo,2 A1 A2 1E+03 1E+02 90 100 1E+05 1E+04
All dimension in microns L = 200 l = 10 wf= wc1 = wp1 = 4 lf = 100 lp1 =6 wc2 = wp2 = 2 lc2 = lp2 = 60

1E+10 1E+09 1E+08 1E+07 1E+06

Total A

Lever 1 Connection Beam Length, micron

Fig. 5.8

(b) Amplification factors, A1, A2 and A, as a function of the T-F connection beam length (SUGAR).

- 105 -

Output Spring Constant, N/m

Amplification Factor, A

Output Spring Constant, N/m

Amplification Factor, A

160 First-kind Levers 140 120 100 80 60 40 20 0 0 10 20 30 40 50 60 70 80 Kvo,2 A1 A2


All dimension in microns L = 200 l = 10 wf= wc1 = wp1 = 4 lf = 100 lc1 =6 wc2 = wp2 = 2 lc2 = lp2 = 60

1E+10 1E+09 1E+08 1E+07 1E+06 1E+05 1E+04 1E+03 1E+02 90 100

Total A Kvo,1

Lever 1 Pivot Beam Length, micron

Fig. 5.8

(c) Amplification factors, A1, A2 and A, as a function of the of lever 1 pivot length (SUGAR).
180
First-kind Levers

160 140
All dimension in microns L= 200 l = 10 lf1 = 100 lc1 = lp1 = 6 wf = wc1 = 4 lc2 = lp2 = 60 wc2 = wp2 = 2 1SS-2SS

Amplification Factor

120 100 80 60 40 20 0 1 2 3 4 5 6 7

Total A Analy. Total A SUGAR A2 A1

1SS-2DS

10

11

12

Width of Pivot Beam, micron

Fig. 5.9 Comparison of SUGAR with analytical results (SUGAR).

- 106 -

Output Spring Constant, N/m

Amplification Factor, A

180
First-kind Levers

160 140
All dimension in microns L = 200 l = 10 lf1 = 100 lc1 = lp1 = 6 wf = wc1 = wp1 =4 lc2 = lp2 = 60 wp2 = 2 Total

Amplification Factor

120 100 80 60 40 20 0 1 2 3 4 5 6 7
A2

A1

10

11

12

Width of 2nd Connection Beam, micron

Fig. 5.10 (a) Effect of second-stage lever connection width (SUGAR).


180
First-kind Levers

160 140
All dimension in microns L = 200 l = 10 lf1 = 100 lc1 = lp1 = 6 wf = wc1 = wp1 =4 lc2 = lp2 = 60 wc2 = 2 Total A

Amplification Factor

120 100 80 60 40 20 0 1 2 3 4 5 6 7
A2

A1

10

11

12

Width of 2nd Pivot Beam, micron

Fig. 5.10 (b) Effect of second-stage lever pivot width on the amplification factors, A1, A2 and A (SUGAR).

- 107 -

160 First-kind Levers 140 120 100 80 60 40 20 0 0 10 20 30 40 50 60 70 80 90 100 A1 A2


All dimension in microns L = 200 l = 10 wf= wc1 = wp1 = 4 lf = 100 lc1 =6 wc2 = wp2 = 2 lp2 = 60

Amplification Factor, A

Total A

Lever 2 Connection Beam Length, micron

Fig. 5.11 (a) Effect of second-stage lever connection beam length (SUGAR)
160 First-kind Levers 140 120 100 80 60 40 20 0 0 10 20 30 40 50 60 70 80 90 100 A1 A2
All dimension in microns L = 200 l = 10 wf= wc1 = wp1 = 4 lf = 100 lc1 =6 wc2 = wp2 = 2 lp2 = 60

Amplification Factor, A

Total A

Lever 2 Pivot Beam Length, micron

Fig. 5.11 (b) Effect of second-stage lever pivot length on the amplification factors, A1, A2 and A (SUGAR).

- 108 -

CHAPTER 6 MULTI-STAGE MICROLEVERAGE MECHANISMS

By stacking multiple single-stage levers together to make a multistage microlever, higher and higher amplification factors could be achieved as more and more single-stage levers are added. As presented in Chapter 5, a two-stage microleverage mechanism can have a much higher amplification factor than a single-stage microleverage mechanism. For instance, in the resonant accelerometer, the amplification factor of the two-stage can be higher than 80 while the amplification factor of a single-stage microleverage mechanism is much lower for a fixed lever ratio. Questions to be asked are: (i) can multiple stages of micro-levers be stacked together to achieve any specified or desired amplification factor? and (ii) is there any limitation for the maximum amplification factor? This chapter presents some answers to these questions related to multistage microlevers. For a specified output system, the achievable amplification factor is not infinite but it is determined by the stiffness of the output system and the lever ratio. The calculation of the maximum achievable amplification factor for a given output system is presented in this chapter. Also discussed is how to obtain the maximum amplification factor by an optimum number of lever stages. The theory for multiple-stage microleverage mechanism design is applied to the resonant accelerometer. Finally, the compliance relationship in the microleverage mechanisms is discussed.

- 109 -

6.1. Analysis of the Multi-stage Microleverage Mechanism


In this section, a general method for calculating the amplification factor of multiple stage microleverage mechanisms is presented. The amplification factor of the i microlever stage can be expressed as a generalized equation as follows:
th

1 Ai =

k vvp , i

( kmo, i + km p, i ) m li Li ( )
(6.1)

1 1 + + + li 2 k vvo, i k vvp, i km o, i k m p, i
th

where kvvo, i is the output vertical (axial) spring constant of the i microlever stage, kvvp,i the vertical (axial) spring constant of the pivot for the i microlever stage, k m o,i the output bending spring constant of the i constant of the pivot for the i
th th th th

microlever stage, k m p,i the bending spring

microlever stage, li the length between the pivot and


th

output system for the i lever arm, Li total length of the i lever arm. While the- sign in the numerator is for the first-kind levers, the + sign is for the second kind. In order to determine the amplification factor of a lever stage in a compound lever, four spring constants in equation (6.1) need to be calculated. The axial and bending spring constants of the pivot in each lever stage are already given in equation (4.23). The calculation of the axial and bending spring constants of the output system of each lever stage follows the same procedure as in the case of the two-stage microlever analyzed in Chapter 5.

- 110 -

The i microlever stage and the connecting beam Ci+1 make a serial connection to become the output system for the downstream lever stage, (i+1) . Therefore, the output axial and rotational spring constants of the (i+1) stage are:
th th

th

k vo, i +1 = 1 1 + kvI , i kvc, i +1

k o, i +1 = 1 1 + k I , i k c, i +1

(6.2)

Among the unknown variables in equation (6.2), the axial and rotational spring constants,

kv v c, i and k m c, i , of the connection beam Ci, can be obtained with equations similar to
(4.23). The input rotational spring constant of the i microlever stage connected together with all the upstream lever stages and the external output system is
th

k m I, i = k mo , i + k m p , i
th

(6.3)

The input axial spring constant of the i microlever stage connected together with all the upstream lever stages and the output system is

k vvI, i =

( kvvo, i + k vvp, i )(k mo, i + kmp, i ) + k vvo, i k vvp, i li 2 ( km o, i + km p, i ) + k vvo, i (Li li ) 2 + k vvp, i Li 2

(6.4)

The procedure of calculating the output axial and bending spring constants of each lever stage is now summarized as follows. Starting with microlever stage #1, the axial and rotational spring constant of a given output system can be readily calculated. Equation (4.23) is used to calculate the axial and rotational spring constants of the pivot beam of lever #1. Equations (6.3) and (6.4) are then used to calculate the input axial and

- 111 -

rotational spring constants of lever #1, while equation (4.23) is used to calculate the axial and rotational spring constants of the connection beam between microlever stages 1 and 2. Equation (6.2) can now be used to calculate the output axial and bending spring constant of microlever stage 2. The calculation process is repeated until the last lever stage connected to an input system (e.g., proof-mass in a resonant accelerometer) is reached. Once all the spring constants in equation (6.1) are calculated, the amplification factor of each microlever stage can be calculated. The total amplification factor is

1 A = A1 A2 Ai An = Ai =
i =1 n n

k vvp, i

( k m o, i + k m p, i ) m li Li ( )
(6.5)

i =1 1 + 1 k m o, i + k m p, i + l 2 i k vvo, i k vvp, i

Similar to equations (5.11a) and (5.11b) for a two-stage microlever, the input spring constant of each lever stage in a n-stage microlever obeys the following equation:

k vvI , i =

k vvI , i 1 k vvI , i 2 k vvo = = ...... = 2 2 (A1 A2 Ai )2 / (e1e2 ei ) A i / e i (Ai 1 Ai ) / (ei 1 ei )

(6.6)

This equation means that, with each microlever stage (with an amplification factor of Ai ) added, the overall input axial spring constant of the compound microlever decreases by a factor of Ai 2 /ei .

- 112 -

6.2 Maximum Amplification Factor of Multistage Lever for a Given Output System
For a given output system with a fixed axial spring constant, the total amplification factor achievable is not infinite, but limited. The maximum achievable amplification factor can be found from the following analysis. Assume the maximum amplification factor Amax has been achieved by n stages of microlevers. From equation (6.6), the overall input spring constant of an n-stage compound microlever, kv v I , n , is

k vvI , n =

(A1 A2 An ) / (e1e2 en )
2

kv vo

kv vo A2 / e

(6.7)

It is seen from equation (6.7) that, as the entire microlever is optimized to achieve greater and greater amplification factor, the overall input spring constant of the n-stage microlever will become smaller and smaller. Adding an (n+1)th microlever stage at the input of an n-stage microlever, the input axial spring constant of the n-stage compound lever becomes the output axial spring constant of the (n+1)th stage microlever, i.e., kv v I , n kv v o , n + 1 (the connection beam is too stiff axially to be considered). When kv v I , n becomes so small that the added (n +1)th stage microlever cannot have an amplification factor greater than 1 within the beam geometry constraints (i.e., minimum beam length and width) set by current micromachining technologies, the n-stage leverage mechanisms may have obtained the maximum possible amplification. In other words, kv v o , n + 1

- 113 -

' ' th kn +1 (Ai 1), where k n +1 is the output axial spring constant of the (n +1) stage microlever when it has an amplification factor of 1. By equaling the amplification factor in equation (6.1) to 1, we have the following equation for a second-kind (n+1)th stage microlever:

' = kn kv vo, n+1 +1

An +1 =1

k mo, n+1 + k mp, n+1


l n+1 Ln+1 l n+12

(6.8)

It is noted that a similar analysis can be done in the case where the (n +1)th stage microlever is a first-kind. In equation (6.8), k mo , n + 1 is a unknown parameter which is dependent on all the upstream microlever stages and the external output system. However, since microlever stages become more and more stiff from the input to the external output system, we have:

km o, n+1 k m c, n+1

(6.9)

From Eq. (6.7) with the total amplification factor reaching a maximum when

Ai + 1 =1, we have:
kv vo
2

kv vI , n A

n +1=1

(A1 A2 An ) / (e1e2 en ) A

kv vo
Amax2 / e

(6.10)

n +1=1

Since kv v c , n + 1 is always much greater than kv v I , n , we have kv v o , n + 1 kv v I , n . Then, equaling the right hand side of equations (6.8) and (6.10) and substituting

k m o , n + 1 with k m c , n + 1 , we have

- 114 -

Amax =

(l

n +1 Ln +1

l n+1

)k

v vo e

k mc, n+1 + k mp, n+1

(6.11)

Taking off the subscript n +1 in equation (6.11) and regarding this added microlever stage as the one which would have the optimized dimensions to give the maximum value on the right hand side of equation (6.11), we have

Amax =

( lL l 2 ) kvvo e
k mc + k mp

(6.12)

From equation (6.12), it can be seen that the maximum amplification factor is dependent on (i) kv v o , the axial spring constant of an external output system; (ii) e, the overall energy efficiency of the compound microlever, and (iii) maximum L and l = L/2 of a lever arm (noting that lL l2 reaches a maximum at l = L/2 for a fixed value of L), (iv) minimum k m c and k m p . It should be noted that the lever ratio ( L/l =2) of the added (n+1)th microlever stage maximizes (lL l2) for the sole purpose of calculating the maximum amplification factor of a given output system, and this has no bearing on the lever ratios of all the lever stages in the compound microlever. The optimum number of microlever stages to achieve Amax depends on the total area available and other design and fabrication issues such as minimum beam length and width. From energy conservation point of view, there will be more energy consumed at the pivot and the connection beams when more microlever stages are added. Therefore, it is ideal to design a compound microlever with the minimum number of lever stages to achieve the desired amplification factor. In the case of using a single microlever to

- 115 -

achieve a high amplification factor, the required lever ratio may be too large and the lever arm too long. Since the microlever arm length is constrained by the design area, multistage compound microlevers will have to be designed to multiply the amplification factor of each individual microlever stage to achieve a higher total amplification factor. The microlever stages in a compound microlever should become stiffer and stiffer from input to output (upstream), or, in another words, softer and softer from output to input (downstream). There are different combinations of microlever stages to allow adjustment in design area and axial spring constants. Although the main trend of the beam dimensional difference between upstream and downstream stages is clear, the calculation of exact dimensions are still tedious and complicated. While all the above analyses are for force amplification, the analysis for displacement will be similar, noting that force and displacement are two inversely-related parameters from the energy conservation point of view.

6.3 Multi-stage Microleverage Mechanism Design in the Resonant Accelerometer


As presented in the previous chapters, a single-stage microleverage mechanism amplifies the inertial force and improves the sensitivity from 3Hz/g to 45Hz/g in a resonant accelerometer. A second-stage microleverage mechanism can have an amplification factor of 80. Would multiple stage microleverage mechanisms in a resonant

- 116 -

accelerometer provide the potential for more force amplification and sensitivity improvement? This will be analyzed in the next sections.

6.3.1 Estimation of maximum achievable amplification factor for the resonant accelerometer
For simplicity in estimating the maximum achievable amplification factor, the efficiency of all the lever stages are assumed to be 1. As the efficiency of compliant leverage mechanism is always less than 1, the calculated maximum amplification factor is always higher than the real maximum value that the leverage mechanism can have. In addition, the calculation of the axial spring constant of the DETF (Double-Ended Tuning Fork) output system has been given in Chapter 4 and the rotational spring constants of connection beam and pivot beam have been given in equation (4.23). Therefore, from equation (6.12), we have

Amax =

k vo l (L l ) = k m c + k m p

12 wf tf l (L l ) / l f wc 3t c / l c + wp 3 t p / l p

(6.13)

For maximum tuning fork beam width wf = 4 m, minimum tuning fork length lf = 100 m, minimum width of pivot and connection beam wp = wc = 2 m, maximum length of pivot and connection beam lp = lc = 60 m (limited by design area available), L = 2l = 210 m, and all beams having the same thickness, the maximum amplification factor is estimated to be 140 from equation (6.13).

- 117 -

Since the energy efficiency is assumed to be 1 for the above estimation, the actual maximum amplification factor achievable should be less than 140. As the maximum amplification factor is 21 for a single stage lever with lever arm ratio L/l of 210/10, a compound lever can be designed to achieve higher amplification factor, i.e., closer to 140. As shown in Fig. 5.6(a), the first-order analytical model predicts a maximum amplification factor of 105 using equation (5.1) or (6.1), at microlever 2 flexure beam width w2 of 2.0 m. The maximum amplification factor of 105 predicted from equation (5.1) or (6.1) is slightly less than that of 140 predicted from equation (6.13), which is largely due to the assumption of 100% lever efficiency (e=1) in deriving equation (6.13).

6.3.2 Feasibility analysis of a three-stage microleverage mechanism in the resonant accelerometer


In order to greatly increase the sensitivity of an accelerometer, multistage microlevers are considered. Since the total available chip area is limited, the dimensions of the proofmass may be decreased by increasing the number of microlever stages. The decrease in the size of the proofmass will reduce the input inertial force and the sensitivity of the accelerometer. Therefore, there is a trade-off between the number of microlever stages and the sizes of the proofmass. The challenge is to determine the optimum lever stage number to achieve the maximum sensitivity. A three-stage microlever for a resonant accelerometer has not been reported in the open literature. By adding another lever stage to a two-stage microlever, the total amplification factor may drop significantly if the added lever stage cannot have an

- 118 -

amplification factor of at least 1. This is not only because a greater portion of the input energy is consumed on bending of the additional flexure beams and a smaller portion is transmitted to output as more microlever stages are added; but also because of the issue of the axial spring constant compliance match between microlever stages. To achieve a desired amplification factor, a compound microlever should be designed with as a fewer number of microlever stages as possible. Difficulty in adding the third stage microlever to achieve a higher amplification factor is not limited to the case of a resonant accelerometer. Many efforts on the design of multiple stage microlever in other applications have remained unsuccessful. The multiple-stage microleverage mechanism in the micro-tweezer designed by Keller (1995) and the displacement amplifier by Bamford et al (1995) at JPL were all limited to twostage microlevers. A compound microleverage mechanism becomes more and more compliant as more stages of the lever are added. As a microlever stage is added at the input, the output axial spring constant of the added lever decreases. Comparing to the output axial spring constant of the adjoining upstream lever stage, the output axial spring constant of the added lever decreases approximately by the square of the amplification factor of the adjoining upstream lever stage. As illustrated in the Chapter 5 Fig. 5.2, in order for the added lever stage to have a force-amplifying effect, its output axial spring constant (or the input axial spring constants of the upstream lever stages) must lie in region II or III. After two single-stage levers are added to form a two-stage lever, the overall mechanism becomes too compliant to allow the addition of another stage with an amplifying effect.

- 119 -

While a two-stage microlever appears to have good compliance match with the original output system, it is difficult to design a third microlever stage with the desired forceamplification effect and reduced compliance. The SUGAR simulation confirms this. The addition of a third-stage lever to the two-stage leverage mechanism in the resonant accelerometer reduces, the output node displacement by two-orders of magnitude. The output system force is proportional to the displacement. Therefore, the output force and the amplification factor are reduced by two-orders of magnitude. A change in the entire mechanism geometry can not bring these up. The compliance issues for the microleverage mechanism will be discussed in more detail in the next section.

6.4 Compliance in Microleverage Mechanisms


In a compound microleverage mechanism design, the compliance of each stage and the overall mechanism become an important issue. As presented in the previous chapters, the upstream stages must be axially stiff and have an input axial spring constant in region II in order for the next downstream stage to have a force-amplification effect. From the viewpoint of energy conservation, the whole mechanism needs to be compliant enough to allow greater input energy to obtain larger output force or displacement. There is a compliance contradiction in the design of compliant mechanisms. The mechanism needs to be compliant enough to allow maximum input energy and axially stiff enough to allow each lever stage to have a force-amplification effect. The mean compliance of a compliant mechanism can be defined by the mutual energy method by Nishiwaki et al (1998). When the applied load is a dummy unit load,

- 120 -

the compliance is the same as the reciprocal of the spring constant. To design a leverage mechanism for maximum force amplification is to find an optimum compliance. It should be stiff enough to allow the next stage or the current stage to have a force-amplification effect and, on the other hand, flexible enough to allow adequate rotation and elastic deformation. The required flexibility and stiffness are conflicting design objectives which are normally encountered in compliant mechanism design (Frecker, 1997). Taking the design of the compound microleverage mechanism in the resonant accelerometer for example, the following analysis further clarifies the point that the flexibility/stiffness of the mechanism needs to be optimally designed to achieve maximum force amplification. The output system can be considered as a spring, so that output force is proportional to the displacement, Fout = kout out . The energy at the output port is:
F out out = k out out 2

(6.14)

In order to have a larger output force, the output energy needs to be maximized. Since the input energy and the output energy are related through the mechanical efficiency, the mechanism needs to be designed to maximize the input energy. For a specified input force, the mechanism needs to be designed to have a low overall stiffness so that the input displacement is large. On the other hand, the lever stages from the input to the output system need to be progressively stiffer axially in order for each lever stage to have a force-amplifying effect.

- 121 -

CHAPTER 7 SOI-MEMS FABRICATED RESONANT ACCELEROMETER


A resonant accelerometer with a two-stage microleverage mechanism of the 1S1D kind was fabricated with the SOI-MEMS process. The microleverage mechanism has a theoretical amplification factor of 80 based on the simulation presented in section 7.3. A trans-resistance amplifier is integrated on the chip to read the shifting frequency and amplifying the signal. This chapter describes the fabrication, testing and analysis of the SOI-MEMS (Silicon-On-Insulator) resonant accelerometer.

7.1 SOI-MEMS Process Run


The fully integrated SOI-MEMS process was developed by Brosnihan (1997, 1998) at UC Berkeley. The process bridges surface and bulk micromachining by integrating 50 micron-thick MEMS structures with the CMOS standard CMOS integrated circuitry. With a Silicon-On-Insulator wafer as the starting material, structures can be etched out of the silicon layer, using the buried bonding oxide as both etch-stop and releasing layer. The thickness and doping level of the device layer can be specified to match the desired structure height and circuit process. In this run, the device layer (doped) has a thickness of 50 microns and a sheet resistance 7.8 /square. The circuitry is processed independently of the MEMS structure by foundry fabrication, which allows

- 122 -

access to better performance and more robust electronics at a lower cost. The CMOS foundry for this run is SMSC (Standard Microsystems Corporation) in Hauppauge, New York. Figure 7.1 is the cross-section of the SOI-MEMS devices consisting of the CMOS circuit region and the MEMS structure region, the isolation trench and interconnects. SOI-MEMS technology is especially suitable for inertial sensor fabrication with the 50 micron-thick MEMS structure as the proof-mass and other mechanical elements. The high-aspect-ratio MEMS structure offered by the process can separate the in-plane and out-of-plane motion mode, reduce cross axis sensitivity and increase the robustness of the accelerometer. By using a single-crystal-silicon structural material, the residual stresses appearing in deposited silicon films due to grain structure is eliminated. Stress induced into the surface of the SOI device layer from back grinding during substrate manufacture is annealed out during the high temperature steps of the CMOS process.

Nitride Isolation Trench Twin-Well CMOS Circuit Region Interconnects MEMS Structure Region Holes for Structure Releasing

N + <100> Si

SiO2 Handle Si Wafer

Fig. 7.1 Cross section schematic of the SOI-MEMS device.

- 123 -

7.2 SOI-MEMS Fabricated Resonant Accelerometer


A resonant accelerometer with a two-stage microleverage mechanism was fabricated with the SOI-MEMS integrated process. The design of the two-stage leverage mechanism is presented in Section 7.3. The SOI-MEMS process integrates 50-micronthick MEMS structure, lateral anchors, with the electronic circuit through the isolation trench. The dielectric-linked, backed-filled, isolation trench separates the circuits and structures regions. Interconnect runs over the top of the trench, connecting the circuitr to the structures. Figure 7.2(a) shows the SEM and Figure 7.2(b) shows the layout of the SOI-MEMS fabricated resonant accelerometer. It consists of four major parts: two Double-Ended-Tuning-Fork resonators as the inertial force gauge, symmetrical two-stage microleverage mechanisms for inertial force amplification, the proof-mass to generate the inertia force under acceleration and the on-chip trans-resistance amplifier to pick-up the sensing current. When under acceleration, the proof-mass will generate an inertial force. Amplified by the microleverage mechanisms, the inertial force will cause an axial loading on the DETF (doubled-ended tuning fork) resonator to shift its resonant frequency. The two identical DETF resonators form a push-pull configuration, providing temperature compensation. Figure 7.3 shows the SEM of the on-chip transresistance amplifier. Figure 7.4 is a schematic of the amplifier circuit designed by Reossig (1998). In order to release the MEMS structure, the 50m structure has uniformly spaced holes as shown in Fig. 7.5. The connection of the lever mechanism to the proofmass is

- 124 -

First stage

Second stage Proof-mass

Trans-R

DETF Resonator

Fig. 7.2 (a)

SEM of the overview of the SOI-MEMS fabricated resonant accelerometer

shown in the center of the SEM photograph in Fig. 7.5. In Fig. 7.6, the lateral anchor in the mechanism is provided by the isolated trench. The accelerometer with 50-micron structure can sustain greater acceleration without buckling compared to the 2-micron structure. With the two-stage microleverage mechanism offering a larger amplification

- 125 -

factor and a thicker proof-mass from the SOI-MEMS run, the SOI-MEMS resonant accelerometer has potential for commercialization.

Figure 7.2(b) Layout Design of the Resonant Accelerometer

- 126 -

Fig. 7.3

SEM of on-chip trans-resistance amplifier (Designed by Roessig, 1998).

Vdd=5v
W=30 L=6 M107 M101 W=6 L=6 M102 W=36 L=6 M1 W=36 L=6 M3 W=60 L=3 M5 W=6 L=6 M11 W=36 L=6 M13 W=60 L=3 M16 W=60 L=3

Vin

M8 W=60 L=3 M7 W=6 L=18

M10 W=60 L=3

Bias M106 W=30 L=3 W=30 M108 L=3 M103 W=30 L=3

M14 W=60 L=3

M17 W=60 L=3

Vout Gain_Adj

M105 W=60 L=3

M104 W=60 L=3 R=6K C0=0.5p Bias_Adj M2 W=36 L=6 M4 W=36 L=6 M6 W=36 L=6 C 0=0.5p M9 W=36 L=6 M11 W=18 L=6 M12 W=36 L=6 M15 W=72 L=6 C0=1p

A. Biasing

B.Amplifier

C. Output Stage

Fig. 7. 4

Schematic of the on-chip trans-resistance amplifier circuit including (A) biasing supply, (B) amplifier, and (C) output stage. (Roessig, 1998).

- 127 -

Fig. 7.5 mass.

SEM micrograph showing the connection of leverage mechanism to proof-

Fig. 7.6

SEM of the lateral anchor

- 128 -

7.3

Two-stage Microleverage Mechanism Optimization


The SOI-MEMS technology only allows the outside anchor provided by the

separation trench, therefore a 1S-1D type two-stage leverage mechanism is the only feasible layout here for the resonant accelerometer. Figure 7.7 shows the schematic of the two-stage micro-leverage mechanism used in the resonant accelerometer. Figure 7.8 shows the SEM photograph of the two-stage mechanism. Both the lever stages are firstkind. The area between the first-stage and the second-stage lever arms is used as a part of the proof-mass. It has an amplification factor of ~80. The dimensions of the mechanism are as follows: DETF length 200 m, width 2 m, connection beam between DETF and the firststage lever length is 10 m, width is 2 m; First-stage pivot length 4 m, width is 2 m, lever arm width 50 m, length from output to pivot is 40 m, from pivot to input is 300 m; Connection beam between first-stage microlever and second-stage microlever is 75 m long and 2 m wide; Second-stage pivot length 200 m, width is 2 m, lever arm width 20 m, length from input to pivot is 330 m, from pivot to output is 15 m. The mechanism has an amplification of 80 from the analytical modeling, which is confirmed by SUGAR simulation. The amplification factors as a function of the flexure beam dimensions are plotted in Figs. 7.9-7.12.

- 129 -

DETF Anchor

Lever 1

Lever 2

Proofmass

Anchor

Anchor DETF

Fig.7.7

Schematic of the two-stage microleverage mechanism in the SOI-MEMS fabricated resonant accelerometer

- 130 -

First stage lever arm

Connection Beam

Proof-mass

Second stage lever arm

First stage lever arm

Proof-mass

Fig. 7.8

SEM of the two-stage microleverage mechanism in the SOI-MEMS resonant accelerometer.

The lengths of the microlever 1 pivot and the T-F connection beam are relatively short compared to those of the microlever 2 pivot and connection beams. The relatively short length for the microlever 1 pivot and connection beams makes microlever 1 stiffer and gives a higher output spring constant for microlever 2, thus ensuring that microlever
- 131 -

2 will have a desirable amplification factor (Su and Yang, 2000). The first-stage lever arm is also wider (50 microns) than the second-stage (25 microns) for the same reason. The use of a relatively long second-stage pivot and connection beam improves the energy efficiency of the second-stage lever and its amplification factor. The amplification factors of the entire leverage mechanism and each individual lever as a function of the width of the tuning fork is shown in Fig.7.9(a). The amplification factors of both microlevers 1 and 2 and the total amplification factor increase continuously with increasing T-F width. This is due to the increase in the axial spring constant of the overall output system and that of microlever 2 output as the T-F width increases. Fig. 7.9 (b) shows the effect of the T-F length on the amplification factors. As the length of T-F beam increases, the amplification factor of both microlever 1 and 2 decreases because of decreasing axial (vertical) output spring constants, kvvo, 1 and kvvo, 2, of both microlevers. The effect of the T-F beam length on the amplification factor is less significant than the effect of beam width. However, the tuning fork width and length are also major factors in influencing the natural frequency of the DETF resonator, which should be taken into account in order to have optimum overall device sensitivity. Figure 7.10 shows the amplification factors as a function of the first stage pivot (a) width and (b) length, respectively. The amplification factors of microlever 1 and the total amplification factor are almost constant at a pivot width of less than 4 m, but decreases rapidly at wider pivot beams, as shown by Fig. 7.10 (a). The rapid decrease in the amplification factor of microlever 1 is mainly due to increased moment bending

- 132 -

energy consumed at the microlever 1 pivot beam as its width increases. Finally, a width of 2 m is chosen for the first-stage pivot in the final design. The sharp decrease of the total amplification factor with increasing microlever 1 pivot width is because the decrease in A1 overrides the increase in A2 (due to increased output spring constant at the output of microlever 2). By comparison, the pivot length ranging from 2 to 20 m has little effect on the amplification factors, as shown in Fig. 7.10 (b). The length of the first stage pivot is finally chosen to be 4 m. The effects of the width and length of the connection beam between the first and the second lever stages on the amplification factors are shown in Fig. 7.11 (a) and 7.11 (b), respectively. The increase in the width of the connection beam 2 causes a sharp decrease in A2 and the total amplification factor A. On the other hand, increasing the length of connection beam 2 leads to rapid increases in A2 and the total amplification factor A. The final design connection beam is 75 m in length and 2 m in width since 2 m is the minimum width by the design rule. The 1st-oder analytical result is shown in the plot for comparison with the SUGAR results. Figure 7.12 shows the amplification factors as a function of the second stage pivot (a) length and (b) width, respectively. The effect of the pivot beam length of microlever 2 on the amplification factors is similar to that of the length of connection beam 2. The amplification factor increases continuously with the increase of the second-stage pivot length. With the design area constraint, the final length of the second-stage pivot length is chosen to be 200 microns. The amplification factor decreases dramatically with increases

- 133 -

in the second-stage pivot width. The second-stage pivot should be as narrow as possible. So should be the second-stage connection beam. The first-order analytical results are slightly higher than the SUGAR simulated results. The difference between SUGAR and analytical results are due to the negligence of the internal horizontal forces in the analytical approach. However, the 1st-order analytical results show the trends of amplification factor change of the first-stage, the second-stage and the entire mechanism due to the geometrical changes. It is found that the DETF length and width changes have greater influences on the amplification factor of the second-stage than that of the first-stage. The first-stage beam width has significant influence on both stages, while the length changes have only a slight effect on any of the amplification factors. The SOI-MEMS run offers 50-micron-thick MEMS structure. The mechanism is more robust and the accelerometer can sustain greater acceleration before buckling. If the structure thickness of the microleverage mechanism and the output system change together, the thickness change has no effect on the amplification factor. The thickness terms are cancelled out in the amplification factor. The actual amplification shows slight derivations from the simulated one due to the holes in the lever arm structure for structure releasing. The structure with holes may have different spring constants compared to the structure without holes. However, there should be no influence on the analytical simulation since the lever arms are assumed rigid in the analysis.

- 134 -

(a)
180
First-kind Levers

160 140
Total A 1st-order Analy. Total A SUGAR Dimension in micron L1 = 300, l1 = 40 lf1 = 200 lc1 = 10, wc1 = 2 lp1 = 4, wp1 = 2 L2 = 330, l2 = 15 lc2 = 75, wc2 = 2 lp2 = 200, wp2 = 2 A2 A1

Amplification Factor

120 100 80 60 40 20 0 0 1 2 3 4 5 6 7

10

Width of T-F Beam, micron

(b)
180
First-kind Levers

160 140

Amplification Factor

120 100 80 60 40 20 0 60
Dimension in micron L1 = 300, l1 = 40 wf1 = 2 lc1 = 10, wc1 = 2 lp1 = 4, wp1 = 2 L2 = 330, l2 = 15 lc2 = 75, wc2 = 2 lp2 = 200, wp2 = 2 Total A 1st-order Analy.

Total A SUGAR

A2 A1

80

100

120

140

160

180

200

220

240

Length of T-F Beam, micron

Fig. 7.9 The amplification factor as a function of (a) the width and (b) the length of T-F.

- 135 -

(a)
180
First-kind Levers

160 140
Total A 1at-order Analy.

Amplification Factor

120 100 80 60 40 20 A2
A1 Total A SUGAR

Dimension in micron L1 = 300, l1 = 40 lf1 = 200, wf1 =2 lc1 = 10, wc1 = 2 lp1 = 4 L2 = 330, l2 = 15 lc2 = 75, wc2 = 2

0 0 1 2 3 4 5 6 7 8 9 10

Width of 1st Pivot Beam, micron

(b)
180
First-kind Levers

160 140

Dimension in micron L1 = 300, l1 = 40 lf1 = 200, wf1 =2

wp1 = 2 L2 = 330, l2 = 15 lc2 = 75, wc2 = 2 lp2 = 200, wp2 = 2 Total A 1st-order Analy.

Amplification Factor

120 100 80 60 40 20 0 2 4 6 8 10 12 14

Total A SUGAR

A2 A1

16

18

20

Length of 1st Pivot Beam, micron

Fig. 7.10 The amplification factors as a function of (a) the width and (b) length of the first-stage pivot.

- 136 -

(a)
180 160 140
Total A 1st-orderAnaly.
First-kind Levers

Amplification Factor

120 100 80 60 40 20 0 0 1 2 3 4 5
A2 A1 Total A SUGAR

Dimension in micron L1 = 300, l1 = 40 lf1 = 200, wf1 =2 lc1 = 10, wc1 = 2 lp1 = 4, wp1 = 2 L2 = 330, l2 = 15 lc2 = 75

10

Width of 2nd Connection Beam, micron

(b)
180
First-kind Levers

160 140

Dimension in micron L1 = 300, l1 = 40 lf1 = 200, wf1 =2

wp1 = 2 L2 = 330, l2 = 15 lc2 = 75, wc2 = 2 lp2 = 200, wp2 = 2 Total A 1st-order Analy.

Amplification Factor

120 100 80 60 40 20 0 10

Total A SUGAR

A2 A1

30

50

70

90

110

130

150

170

190

Length of 2nd Connection Beam, micron

Fig. 7.11 The amplification factors as a function of (a) the width and (b) length of the connection beam 2 between the two lever stages.

- 137 -

(a)
180
First-kind Levers

160 140
Total A 1st-order Analy.

Amplification Factor

120 100 80 60 40 20 A2 0 0
A1 Total A SUGAR

Dimension in micron L1 = 300, l1 = 40 lf1 = 200, wf1 =2 lc1 = 10, wc1 = 2 lp1 = 4, wp1 = 2 L2 = 330, l2 = 15 lc2 = 75, wc2 = 2

10

Width of 2nd Pivot Beam, micron

(b)
180
First-kind Levers

160 140

Dimension in micron L1 = 300, l1 = 40 lf1 = 200, wf1 =2

wp1 = 2 L2 = 330, l2 = 15 lc2 = 75, wc2 = 2 wp2 = 2 Total A 1st-order Analy. Total A SUGAR

Amplification Factor

120 100 80 60 40 20 0 20

A2 A1

60

100

140

180

220

260

300

340

380

Length of 2nd Pivot Beam, micron

Fig. 7.12 The amplification factors as a function of (a) the width and (b) length of the second stage pivot.

- 138 -

7.4 DETF Natural Frequency


The SOI-MEMS fabricated DETF resonator has already been shown in Fig. 7.2. For the resonant accelerometer, although the other end of the resonator is connected to the flexible leverage mechanism instead of being anchored to the substrate, it is still considered as a fixed end in the following analysis of beam vibration frequencies. The displacement perpendicular to the beam and the rotation angle at its joint with the proofmass are zero, fulfilling the boundary condition of a fixed end beam. The frequency of the resonator can be considered as a fixed-fixed beam with the comb-drive structure as center-attached mass for simplicity. As a matter of fact, the comb structure provides a uniformly-distributed lateral force when a DC voltage is applied. The electrical force effect can be considered as a electrical spring which softens the mechanical spring constant, therefore decreases the resonant frequency. The thickness does not effect the resonant frequency. There are mainly two approaches to calculate the resonant frequency: the classical approach and the energy approach. While the classical approach calculates all frequencies and vibration modes, the energy method is used to approximately estimate the fundamental frequency. In the case of DETF in resonant sensor applications, the fundamental vibration mode of the DETF corresponding to the lowest natural frequency is actuated. The energy method or Rayleighs method is commonly used to calculate the natural frequency (Howe

- 139 -

Driving

DETF

Comb-Drive

Sensing

Fig 7.13

SEM of the SOI-MEMS fabricated DETF resonator.

1987). A suitable function is assumed for the deflection which satisfies the boundary condition, and then the maximum strain and kinetic energies are calculated and equated to solve for the approximate values of the fundamental natural frequency. For conservation of energy, the strain energy in the position of maximum deflection must equal to the kinetic energy when passing through the equilibrium position. The accuracy

- 140 -

of the solution of the natural frequency depends on how close the assumed shape function is to the actual vibration shape. The energy method can estimate most of the fundamental frequencies when exact solutions from the classical method are difficult to pursue. For the DETF resonator in the micro-accelerometer with comb-drive actuation and sensing structure connected as lumped mass, Roessig (1998) presented in his Ph.D. dissertation both the classic beam vibration method and energy method for the natural frequency calculation. For a fixed-fixed beam without attached mass, the exact solution of the basic lateral vibration frequency can be found in textbooks (Harris and Crede, 1976).

0 =

22.4 l2

EI

(7.1)

where I = tw3/12 is the moment of inertia, t beam thickness, w beam width, density, and S cross-section area (S = tw). Substituting these variables to equation (7.1), we have:

0 =

6.47 w E l2
3

(7.2)

Using E = 150 GPa, = 2330 Kg/m , we have the natural frequency expressed as:

0 =

51912 w l2
-6 2

(7.3)

Assuming w = 2x10 m, l = 200x10 m , then f = /2 = 8266 w/l . The natural


-6

frequency is calculated to be 414 KHz which agrees with 412KHz from Roessigs

- 141 -

calculation (Roessig, 1998) with the classic method. The effect of the DETF length and width on the natural frequency can be seen directly from this formula. It is useful for optimizing the geometry of the resonator for the maximum amplification factor of the leverage mechanism and improving the measurement sensitivity. The natural frequency of a fixed-fixed beam depends on the beam length and width, but is independent of the beam thickness. The actual frequency of the resonator is less than the above-calculated value because the comb-drive structure mass is ignored in the calculation. The effect of comb-drive structure mass is analyzed in the following. The natural frequency of a fixed-fixed beam with concentrated center mass M can be found through Rayleighs method as follows (Roessig, 1998).

0 =

192 EI l 3 ( M + 0.4 wlt )

(7.4)

It is noted that equation (7.4) reduces to (7.1) when M = 0 (without attached mass). The natural frequency in equation (7.4) can be approximated by a lumped parameter method as follows.

0 =

K M + 0 .4 m

(7.5)

where K = 192 EI /l and K is the spring constant for a fixed-fixed beam when loaded at the center. Meff = M + 0.4m is called the effective mass where M = at and m = at, a is the surface area of the comb drive structure, a surface area of the resonator (a = lw), t

- 142 -

thickness of the structure, and I = tw /12 is the moment of inertia. Therefore, equation (7.4) can be rewritten as:

0 =

16 EW 3 3 L (a'+0.4a)

(7.6)

If the comb drive and the resonator structure have the same thickness, the natural frequency of the resonator does not depend on the structure thickness, but on the total surface area, the length and width of the resonator beams, Youngs modules and the density. Using E = 165 GPa, w = 2x10 m, l = 200x10 m, = 2330 Kg/m3, and comb-6 -6

drive area being approximately 4 times the area of the tuning fork, the natural frequency is calculated to be 130KHz. An alternative way of calculating the frequency of the DETF is using ABAQUS or SUGAR simulations. SUGAR is used to simulate the DETF resonator with the center comb-drive mass. Figure 7.14 shows the SUGAR simulated structure with the combdrive as attached center mass. The netlist file is attached in Appendix F. The frequency is found to be 129KHz which agrees with the foregoing calculated value of 130KHz. When DC voltage is applied to the driving pad and the DETF beam, there is a distributed electrical static force acting laterally on the beam which soften the lateral mechanical spring constant. With the high aspect ratio structure, the electrical softening effect cannot be ignored due to the large electrical static force generated at the high capacitance plate [Boser, 1997]. The effect of the electrical spring constant softening can be expressed as follows:

- 143 -

x 10m 0

-4

SUGAR - DC ANALYSIS DISPLAY 2

-0.5

-1 3 6

-1.5 8

-2

4 -1 -0.5 0 0.5 1 1.5 x 10m


-4

Fig. 7.14 SUGAR modeling of the DETF resonator with comb-drive mass.

x 10m 0

-4

SUGAR - DC ANALYSIS DISPLAY 2

23 21 -0.5 7

-1

33 31 3 34 32

-1.5 8 44 41 -2

4 -1 -0.5 0 0.5 1 1.5 x 10m


-4

Fig. 7.15 SUGAR modeling of electrical tuning of DETF resonator frequency.

- 144 -

K elec =

2 AV 2 g3

(7.7)

where A is the area of the capacitor, g gap and V applied voltage. The electrical softening effect on the resonant frequency can be seen by subtracting the electrical spring constant from the mechanical spring constant. SUGAR can also be used to simulate the electrical softening effect. Figure 7.15 schematically shows the node information for the gap. The netlist files for simulation of the DETF resonator are attached in Appendix F and G.

7.5 Sensitivity Analysis of the Resonant Accelerometer


The high aspect ratio (50 microns thick) structure from the SOI-MEMS technology is valuable for inertia sensor applications. It increases the robustness of the sensor, e.g., the resonant accelerometer can sustain higher acceleration without buckling. In most of the other applications, the high-aspect-ratio proofmass also increases the sensitivity. For resonant accelerometers, the sensitivity in hz per g will not change as the thickness changes, but will increase due to a frequency decrease caused by the electrical softening effect. The natural frequency of a fixed-fixed beam under an axial load F can be found by the energy method:

= 0

FL2 1 40 EI
- 145 -

(7.8)

where 0 is the natural frequency without axial force F. The +sign is for tension and sign for compression (noting that the coefficient is 5/14 for a beam with one fixed end and one free end; 1/ for a beam with two hinged ends; 1/40 for a beam with double fixed ends). Taking the derivative of with respect to F, the sensitivity can be expressed as [Roessig, 1995].
1 M lf eff 0
2

df 1.21 = dF (2 ) 2

(7.9)

The axial force is the inertial force from the proof-mass when under acceleration, F = Mg, and M = LmWmt is the proof-mass (Lm and Wm are proofmass length and width, respectively). The sensitivity with respect to the acceleration is not dependent on the thickness as the thickness term is cancelled out.
M M lf eff 0

df 1.21 = dg (2 ) 2

(7.10)

The sensitivity is a function of the fundamental natural frequency, the resonator geometry as well as the ratio of the surface area of the proof-mass to the surface area of the resonators. The actual sensitivity is the calculated value multiplied by the amplification factor of the microleverage mechanism. Using 130KHz as the basic frequency and proof-mass area 750 x 450 + 2 x 80 x 750 m2, the sensitivity is calculated to be 2Hz/g. With the two-stage microleverage mechanism having a force-amplification factor of 80, the sensitivity of SOI-MEMS accelerometer is improved to 160Hz/g. The electrical lateral force decreases the frequency and increases the sensitivity.
- 146 -

7.6 Optimization of DETF Resonator


The amplification factor of a two-stage microleverage mechanism depends on the geometry of the DETF resonator which is the output system of the mechanism. As shown in Figure 7.16 and 7.17, the amplification factor increases with the DETF width and decreases with the DETF length. From the force-amplification point of view, the DETF resonator should be as wide and as long as possible. However, the geometry of the DETF also influences the resonator frequency and the sensitivity. As seen from equation (7.3), the natural frequency of the DETF resonator increases with the DETF width and decreases with length. The natural frequency also influences the sensitivity of the resonator as shown in equation (7.10). The sensitivity of the resonator increases with the DETF length and decreases with the width. The overall sensitivity of the accelerometer is the resonator sensitivity multiplied by the amplification factor of the mechanism. Therefore, there is an optimum DETF geometry where the overall sensitivity of the accelerometer is at the maximum. As shown in Fig. 7.16, the overall sensitivity first increases with increasing DETF width up to 1.5 microns at which the sensitivity is at the maximum. When the width is greater than 1.5 micron, the sensitivity decreases. Increases in the DETF length cause increases in the sensitivity as shown in Figure 7.17, although the leverage mechanism amplification factor decreases. The final geometry of the DETF resonator is chosen to be 2 m wide and 200 m long, and the overall sensitivity is 150Hz/g.

- 147 -

Fig. 7. 16 Sensitivity as a function of the DETF beam width.

Fig. 7. 17 Sensitivity as a function of the DETF beam length for a series of beam width
- 148 -

7.7 Testing of the SOI-MEMS Fabricated RXL


Figure 7.18 lists the process flow after the SOI wafer is back from the foundry. Post-processing of the SOI wafer includes releasing of the MEMS structure with HF etching and cutting the wafer to dies. Both front- and back-side releasing are used. The die then needs to be wire-bonded and packaged with the West-bond machine in the UC Berkeley Microfabricaiton laboratory. Then, the accelerometer is ready for testing.

HF

SOI Wafer
Releasing

SOI Die

Westbond Packaging

SOI Chip

Off-Chip Circuit

PC-Board

MMR

Test Result

Fig. 7.18 SOI-MEMS RXL post-process steps.

7.7.1 Experimental testing set-ups


Figure 7.19 shows the entire testing apparatus which includes the on-chip part, off-chip circuit and the oscilloscope. The on-chip part includes the DETF resonator and an on-chip trans-resistance amplifier. The off-chip circuit is a unity gain feedback buffer as shown in Fig. 7.20. The operational amplifier is a LM318N/218/118 model with 8 pins. The DETF resonator in the accelerometer has a balanced excitation scheme [Roessig 1995]. The inside comb drive pad close to the fork is for driving while the outside comb drive is for sensing. The on-chip trans-resistance amplifier SEM and schematic are shown in Fig. 7.3 and Fig. 7.4, respectively.
- 149 -

Driving

Sensing Current Trans-R

Voltage Output

LM318 Unit gain feedback buffer

Oscilloscope

DETF

Comb Drive Off-Chip

On-Chip

Fig 7.19

Testing apparatus for the SOI-MEMS resonant accelerometer.

10K

10pf +15V Vout from Chip 10K LM318 + Vss from Chip Ground -15V Ground 100 Output to Oscilloscope

Fig. 7. 20 Off-chip unit gain buffer schematic

7.7.2 Testing of the DETF resonator


The first prototype testing was performed in air on a probe station. The probe station has five probe-tip manipulators and a video camera. The released die was placed on the platform with the probe-tip touching driving, sensing and output pads and also power and ground pads. The electrical apparatus used were: HP E3612A power supply, HP4195A spectrum analyzer, a function generator for the AC power supply and an
- 150 -

oscilloscope. Two DC power supplies with maximum DC voltage of 100 volts were also used, one for the driving voltage, the other for the sensing voltage. First, DC and AC voltages were applied to the two driving pads. The DC and AC voltages were in series. The AC voltage had a 10 Volt DC supply with a 5 Volt offset. The resonators from front- and back-side releasing methods have different performances. A typical backside released resonator starts to vibrate at resonant frequency 60-80KHz when the AC is at 3 Volts. The structure looks vague while vibrating. Since it is tested in air, the air damping causes a low Q-factor. The frequency bandwidth is wide and the vibration amplitude is small. However, when the frequency drops to 10KHz or much lower, such as 80Hz, the resonator also vibrates with large amplitude. Typically, resonators start to move when the AC is at 60Khz and the DC is less than 50 volts. Most of the pads started to burn when the voltage reaches 50 Volts. Thus the DC voltage needs to be lower than 50 Volts. The DC voltage needed for the SOI-MEMS resonator is much lower than typical DC driving voltage for a comb-drive actuated resonator. This is possibly due to the 50-micron-thick structure from the SOI-MEMS process. The resonator beam and the back of the comb-drive formed a parallel-plate capacitance actuation system. The gap was about 6 m. The parallel-plate actuation force was much larger than the comb-drive force. The actuation force generated due to fringing fields was smaller and also decreased by the over-etching post-process. The resonator is in fact actuated through the parallel-plate capacitance. The parallel-plate capacitance actuation acted like a negative spring constant which tuned the natural frequency of the resonator to a lower frequency. The backside released resonator structures seemed to

- 151 -

work better than the front-side released ones, but the circuits did not function well. The frequency needs to be found through other detection methods, such as off-chip transresistance amplifier or by observation of the vibration amplitude from a videotape. The circuit for the front-side released chip worked, but the vibration magnitude of the DETF resonator was much smaller than that of the backside released ones.

After testing the resonator, three other probe-tips were put on the top of Vdd, Vss and Vout pads. A 5-volt DC supply was connected to the Vdd, the Vss pad was grounded, and the output was connected to the oscilloscope. The resonant frequency was also found by measuring the frequency of the output sine-wave signal. Figure 7.21 shows the magnitude and the phase when the DC voltage was 1 Volt.

7.7.3 Testing of the accelerometer


After testing the resonator, the die was packaged with the West-bond machine with both 14-pin and 40-pin packages. The package was then wire-connected to the unity gain off-chip circuit. The entire PC board was first dropped down freely to have acceleration g. The frequency shift matched with the predicted sensitivity. The two-stage micro-leverage mechanism improves the sensitivity greatly as designed. The SOI-MEMS fabricated resonator had larger vibration amplitude. The testing was performed at one fixed frequency.

- 152 -

-12 log10(magnitude) -14 -16 -18 -20 2 10 0 phase(degree) -50 -100 -150 -200 2 10

10

10 Frequency (Hz)

10

10

10

10 Frequency (Hz)

10

10

Fig. 7. 21 Plots of magnitude and phase of the DETF resonator. The disadvantage of testing in air is the low Q-factor due to air damping. The devices both in die and package were also tested in the MMR vacuum chamber. For the die testing, due to limited spacing at the chamber, the microscope did not have a high enough magnification to allow one to see the vibration. Electrical detection is not possible since there was only four probe-tip manipulators available. The resonant accelerometer is very attractive for applications in car and navigation systems. With the two-stage leverage mechanism to amplify the inertial force, the sensitivity is great increased. The sensor has good potential for commercialization.

- 153 -

7.8 Future Work on Resonant Accelerometer


The SOI-MEMS fabricated resonant accelerometer with a two-stage

microleverage mechanism can have a sensitivity of ~160Hz/g (2Hz/g multiplied by 80 which is the amplification factor of the leverage mechanism). The DETF resonator functions better in vacuum which eliminates air damping and allows a high Q factor. Future work includes sealing the entire resonator and the proof-mass with a cap using metal bonding (Cohn, 1997). The SOI-MEMS process offers a SiO2 passivation layer and exposes the metal, which makes it convenient to add a cover cap. It should be laid out with the metal pads outside the perimeter of the structure. The difficult task is to make a big enough cover cap to cover the entire resonator and the proof-mass without warping. Other possibilities include sealing the entire die with vacuum packaging. Further work on the resonant accelerometer also includes the design of the twoaxis accelerometer with two-stage microleverage mechanisms, using either the SOIMEMS run or the Sandia processes as shown in Fig. 7.22. The Sandia integrated process also provides on-chip integrated circuitry. The second stage needs to have the pivot and output (i.e., connection beam) on same side of the lever arm in order to fit four levers and resonators all together.

- 154 -

(A)

(b)

Fig. 7.2Schematic of a two-axis resonant accelerometer with two-stage microleverage mechanisms: (a) 1S-2S type, (b) 2S-2S type.

- 155 -

CHAPTER 8 EXPERIMENTAL VERIFICATION WITH MACRO MODEL

The microleverage mechanism in the SOI-MEMS fabricated resonant accelerometer has only one configuration and cannot be modified for experimental validation of other configurations unless a totally separate device is fabricated. In addition, the SOI-MEMS fabrication process suffered from a low yield. In order to verify the effect of different configurations on the amplification factor of a leverage mechanism, a scaled-up macro model was built with aluminum sheet and plate. This chapter presents the building of the macro model and the experimental results.

8.1 Building the Macro Aluminum Model


There are two main parts in the aluminum model: the backing plate and the twostage leverage mechanism. Figure 8.1 shows the backing plate with parallel slots machined for anchoring the pivot beams and the output beams. The long slots allow easy adjustment of beam length and change of lever kinds, e.g., from first-kind to second-kind. The plate is made of aluminum alloy 6061-T6. Figure 8.2 shows the entire two-stage leverage mechanism with detailed dimensions.

- 156 -

1.0" 1.0" 1.0"

12.0" 10.0" 11.0" 24.0" 10.0" 11.0"

Symmetrical 6061 Aluminum Backing Plate, 0.250" Thick All Slots 1/4" Wide, Shape of Slot Ends not Critical

Fig. 8.1

Aluminum backing plate for anchoring the leverage mechanism.

- 157 -

1.0" 1.0" 1.0"

12.0" 10.0" 11.0" 24.0" 10.0" 11.0"

Symmetrical 6061 Aluminum Backing Plate, 0.250" Thick All lever arms and pivot anchor pieces are made of 6061 hollow square tube. All pivot beams and connection beams are made of 7475 sheet strips.

Fig. 8.2

Scaled-up two-stage leverage mechanism made of aluminum.

- 158 -

Load Cell Sensor Aluminum Plate

Anchor 1st-stage Pivot 1st-Stage Lever

Connection Beam 2nd-Stage Lever Arm

Weight Input

Fig. 8.3

A photograph of the macro aluminum leverage mechanism.

A photograph of the macro-scale two-stage aluminum lever is shown in Fig. 8.3. The S-shaped load cell is used to measure the amplified force. The output system is two parallel beams, similar to the DETF resonator. The two-stage compliant lever shown in Fig. 8.3 is a 1S-2S type. The lever arms and pieces for anchoring pivot beams are made of a hollow square tube of aluminum alloy 6061-T6 with a wall thickness of 3.35 mm and outside dimensions of 25.4 mm x 25.4 mm. All the pivot and connection beams are made of cold rolled sheet strips of aluminum alloy 7475, with their ends bolted to either the lever arms or the anchor pieces fixed to the backing plate. The input forces are applied to the second stage lever by hanging weights at ends of the lever arms.

- 159 -

The flexure beams of the second stage lever are narrower and longer than those in the first stage to improve the amplification factor. The exact dimensions of the two-stage aluminum leverage mechanism are as follows: thickness of the entire structure t = 19 mm, output beam width wo = 1.8 mm and length lo = 85 mm, first-stage lever arm distance between input and pivot, L1 = 221 mm, between connection beam and pivot l1 = 23 mm, width of lever 1 connection beam and pivot beam, wc1 = wp1 = 1.8 mm, length of lever 1 connection beam and pivot beam, lc1 = lp1 = 22 mm, second-stage lever arm distance between input and pivot, L2 = 257 mm, between connection beam and pivot l2 = 19 mm, width of the lever 2 connection beam and pivot beam, wc2 = wp2 = 1.04 mm, length of lever 2 connection beam lc2 = 85 mm, and pivot beam lp2 = 136 mm,

8.2 Experimental Testing with the Macro Model


Figure 8.4 shows schematically the experimental set-up. The input force is generated by the weight at the input of the mechanism. A load cell is placed at the output to measure the output force. A bridge is connected to the load cell to amplify the signal from the load cell to a voltage output. The voltage output signal then goes through an A/D converter card in the computer and LabTech software is used to read the final voltage output. Alternatively, a voltmeter can be used to read the voltage output.
- 160 -

The displacement at the input caused by the lever arm weight was over 10 mm when the structure was standing vertically without any input force (i.e., no weight hung from lever arm 2). Under such large displacement, the classic elastic-beam-bending theory may not give accurate beam deflections. To minimize the effect of lever arm weight, the model could be placed horizontally and the input forces applied through two pulleys. With the lever structure standing vertically, experiments were carried out to measure the force amplification factors. First, the load cell was calibrated with a series of weights and a linear relationship was established between the weight and voltage difference before and after the addition of a weight. It was found that a 1 lb weight is equivalent to 0.148v recorded on the voltmeter. Then, with the entire leverage

Load-cell Sensor Bridge LabTech Data Acquisition Software

A/D

Computer

Macro Model

Fig. 8.4 Experimental set-up for macro-model verification. mechanism connected to the load cell, the voltage readings were taken at 0 lb (before loading), during loading, and 0 lb (after unloading). This process was repeated until

- 161 -

consistent voltage readings were obtained. A series of weights, , , and 1 lb, were symmetrically applied to the input ends of two second-stage lever arms. Two lever configurations, one 1S-2S type and one 1S-2D type, were tested and the results are shown in Fig. 8.5. The 1S-2S configuration had a force amplification factor of ~50 with a slight decrease at heavier weights. In comparison, the 1S-2D type had a force-amplification factor of ~22, but with a slight increase at heavier weights. These measured amplification factors are much lower than the ideal amplification factor of 120 as calculated below: L1 = 221 mm, l1 = 22.8 mm, A1 = L1 / l1 = 9.67, L2 = 257 mm, l2 = 19 mm, A2 = L2 / l2 = 13.5, the ideal total amplification factor A = A1A2 = 130. The analytical equation (4.8) predicts a value of 90. The amplification factor of the first stage, A1, is calculated by first substituting beam dimensions and E = 70GPa (for aluminum alloy) into equation (4.23) to calculate all the spring constants (e.g, kvvp,1 = EWp,1 t / lp,1 = 6.9 x 102 N/m) and then substituting the calculated spring constants into equation (4.8). The amplification factor of the second stage, A2, is calculated by first substituting beam dimensions and E = 70GPa (for aluminum alloy) into equations (5.4) to (5.7) to calculate all the spring constants and then substituting the calculated spring constants into equation (4.8). With respect to SUGAR simulation of 1S-2S type, the netlist file is given in Attachment I and the node information is the same as in Figure 5.5(c). The amplification factor by SUGAR simulation is 80. For the 1S-2D type, the netlist file is given in Attachment H and the node information is the same as in Figure 5.5 (b). The

- 162 -

amplification factor of 1S-2D type by SUGAR simulation becomes 63, which is slightly lower than that of the 1S-2S type. The experiment results shown in Fig.8.5 has a lower amplification factor than those predicted by the first order analytical results as above (using equation 4.8). This is qualitatively similar to the SUGAR results which also show a lower amplification factor than the 1st order analytical results (e.g., Fig. 5.9). However, the experimentally measured difference in the amplification factor between 1S-2S and 1S-2D macrolevers was much greater than that predicted by SUGAR simulation. For the dimension used at the macro-level experiment, the quantitative difference is much larger. Although there is a negligible variation with the input load, the structure weight contributes to a great amount of the elastic deformation as described in the next paragraph. In all, the measured amplification factors of the macro-levers are much lower than the predicted values, and the difference in the measured amplification factors between the 1S-2S and 1S-2D types is much greater than that predicted by SUGAR. All these are most likely due to the large displacements (even under lever arm gravity only, without any input force) at the second-stage lever arm input. As the dimensions of a microlever are scaled up with a scaling factor R to a macrolever, the surface area of the macrolever arm is enlarged by a factor equaling the square of R. Consequently, the weight of the lever arm is scaled as the cubic of R and may no longer be ignored as in the micro-scale. If the macrolever is positioned horizontally, the weight of the lever arm may bring the structure into contact with the substrate. On the other hand, if the macrolever is positioned vertically, the weight of the lever arm which is in the same direction as the input force may by itself cause significant displacement of the structure and hence affect

- 163 -

the amplification of the input force. Since the weight is also shrunk while at micro level, making device at micro-scale has many advantages verse the same device in macro level. Fig. 8.5 Measured amplification factors as a function of the input load for 1S-2S and 1S-2D two-stage aluminum levers.

80

Measured Amplification Factor

1S-2S

1S-2D

60

40

20

0 0.5

1.0

1.5

2.0

Total Input Load (weights), lbs

- 164 -

CHAPTER 9 MICROLEVERAGE MECHANISM FOR DISPLACEMENT AMPLIFICATION

This thesis provides a fundamental study of the microleverage mechanism design for both force and displacement amplification. The foregoing chapters present the design theory of microleverage mechanisms and lay an important foundation for force amplification. Compliant microleverage mechanisms have potentially wide applications in micro-electro-mechanical systems, such as inertial sensors, including gyro and accelerometer, micro-valve (William, 1999), micro-tweezers (Keller, 1995), disk-drive suspension for displacement amplification (Chen and Horowitz 2000). The foregoing chapters present the analysis and synthesis of the compliant leverage mechanism mainly for force amplification. Displacement and force are two reciprocal factors. However, design of leverage mechanism for displacement amplification differs from that for force amplification in several aspects. The following sections present the analysis of displacement-amplification leverage mechanisms and its application in several practical design cases.

9.1 Different Configurations of Leverage Mechanism Displacement Amplification


- 165 -

for

For single-stage leverage mechanisms, there are two kinds of levers for displacement amplification. As described in Chapter 2 section 2.3 for compliant leverage mechanism classification, a third-kind lever (input lies between the pivot and output) can be used for displacement amplification. A first-kind lever (pivot lies between the input and output) can be used to amplify either force or displacement depending on which distance is greater between the two: (i) the distance between the pivot and the input and (ii) the distance between the pivot and the output. Figure 9.1 schematically shows the two kinds of single-stage leverage mechanisms for displacement amplification. Stacking two lever stages together forms two-stage leverage mechanisms. Building on the classification scheme developed in Chapter 2 and applied to the twostage leverage mechanism for displacement amplification, there are sixteen different configurations formed by the two kinds (first and third kind) of single-stage levers and different subgroups according to whether the pivot and output are on the same (S) or different (D) sides of the lever arm. Table 9.1 shows these sixteen configurations of twostage leverage mechanisms for displacement amplification. In the two-stage leverage mechanism for displacement amplification, the lever stage connected to the input is still called the first stage, the one connected to the output is called the second-stage since displacement and force are reciprocal. Depending on the different design constrains in the actual application case, a specific configuration combined by different configurations can be used.

- 166 -

(1) First-kind Lever for Displacement Amplification

Pivot

Output

L Input

(2) Third-kind Lever for Displacement Amplification

Output l Pivot L

Input

Fig. 9.1 Two kinds of leverage mechanisms for displacement amplification.

- 167 -

Table 9.1. Different Configurations of Two-stage Microleverage Mechanism for Displacement amplification NO. 1 CODE TWO-STAGE LEVER
CONFIGURATION

NO.

CODE

TWO-STAGE LEVER
CONFIGURATION

1D-1D

3D-1D

1D-3D

10

3D-3D

1D-1S

11

3D-1S

1D-3S

12

3D-3S

1S-1D

13

3S-1D

1S-3D

14

3S-3D

1S-1S

15

3S-1S

1S-3S

16

3S-3S

- 168 -

9.2 Amplification Factor of Leverage Mechanism for Displacement Amplification


The leverage mechanism for displacement amplification differs from that for force amplification in several aspects. In many application cases of the leverage mechanism for displacement amplification, the input is normally connected to a certain actuation structure, such as a thermal actuator, magnetic actuator or electrostatic actuator. The output end is free standing or connected to some output system (carrying some device such as a gambol in the disk-drive suspension application case). If the lever arm is considered as a rigid body, each point on the lever arm has the same rotation angle. The ratio of the output displacement to the input displacement, the displacement amplification factor or the geometry advantage is simply the ratio of the distance of the output and input to the pivot. The displacement amplification factor for a free standing output system is more straightforward comparing to the force amplification factor. The output displacement equals the input displacement times the lever ratio l/L where l is the distance from output to the pivot and L is the distance between input and the pivot. The input displacement is determined by the input force and the input port compliance of the entire mechanism. If the output end of the mechanism is not freestanding but connected to other output systems, such as other flexures, then the amplification factor becomes more complicated and a more thorough analysis must be performed. The amplification factor

- 169 -

also depends on the output system compliance, as in the case of a microtweezer picking up an object. The amplification factor also varies according to the object stiffness. Displacement is easier to measure as compared with force at the micro-scale. It can be accurately measured by a Vernier gauge. Testing structures can be built to verify the design theory and study the change of the amplification factor when the mechanism is connected to different output systems of different compliance. Figure 9.2 is a test structure of a two-stage microleverage mechanism for displacement amplification. Thermal expansion actuators based on both length change and temperature difference can be used to generate an input displacement. Multiple stages of microleverage mechanisms can be designed to amplify the input displacement. The test structure can be fabricated by the MCNC MUMPs (Multi-User MEMS Process) run and the heatuator layout is available from the Consolidated Micromechanical Element Library (CaMEL) (http: //www.memsrus.com/cronos/svcscml.html). There are many issues associated with the displacement amplification, such as compliance requirements of different lever stages, geometry optimization, and the combination of different kinds of micro-levers. The amplification factor can be compared between the single-stage and two-stage while the input displacement is the same. In order for the two-stage leverage mechanism to have an amplification effect, it may need a greater input force since adding more stages increases the compliance of the mechanism. To generate the same amount of input displacement, single-stage and two-stage mechanisms need different input forces since they have different compliance at the input port.

- 170 -

Similar to the force amplification, the displacement amplification factor also depends on the geometry, such as the connection beam length and the width between the first- and the second-stage lever. The lever arm needs to be rigid. The geometry of the first-stage and second-stage pivot also influences the amplification factor. The connection beam between the input and the mechanism plays an important role in influencing the amplification factor. The mechanism functions differently depending on whether the output end is freestanding, or connected to other output system. A flexure beam may be added as the output system to study the influence of the output system compliance on the amplification factor.

Fig. 9.2 Test structure of the two-stage microleverage mechanism for displacement amplification. Leverage mechanisms for displacement amplification have been used in many applications, such as disk drive suspension (Chen and Horowitz 2000), microtweezer(Keller, 1995), microvolve (William, 1999) etc. In most cases, only the singlestage leverage mechanisms are used. The following sections present the analysis and design of two-stage or multiple-stage leverage mechanism for displacement amplification in several application cases.
- 171 -

9.3 Displacement Amplification Leverage Mechanism in a Silicon Disk-Drive Suspension


The leverage mechanism used in a PZT-actuated silicon suspension for displacement amplification is shown in Fig. 9.3 (a) and (b) [Chen and Horowitz 2000]. The silicon suspension is used in Magnetic Hard Disk Drives (HDDs). The suspension integrates a gimbal structure at the tip and is actuated by PZT strips. When a voltage is applied, the PZT thin films generate compression (push) at one side and tension (pull) at the other side. This push-pull motion rotates the arm structure, which is anchored in the center to the fixed part by the pivot, a small angle. The arm structure is a one-piece rigid body which carries the gimbal at the end. The leverage mechanism is used to transform the axial push-pull movement of the PZT strips into an in-plane angular motion of the gimbal structure, with the amplification factor defined as the ratio of inplane gimbal displacement (~1 m is desirable) in the direction perpendicular to PZT strips to the axial displacement of the PZT strips. The leverage mechanism has an amplification factor of 3 which is the ratio of the distance between the strip and pivot (at center) to the distance between pivot and gimbal tip. For the suspension design, it is better for the strip to be as close to the gimbal as possible. The amplification factor needs to be increased without increasing the distance of the strip to the tip. Figure 9.4 shows a proposed symmetrical two-stage leverage mechanism design that brings the PZT strip closer to the gimbal with a greater displacement amplification factor. The geometry is further analyzed and optimized with SUGAR for an amplification factor of 30.

- 172 -

Fig. 9.3

(a) PZT-actuated silicon suspension (Chen and Horowitz).

Fig. 9.3

(b) Schematic of the silicon suspension design (Chen and Horowitz).

1st stage lever

2nd stage lever

pivot

PZT Strip

Fig. 9.4 Two-stage leverage mechanism design in the suspension.

- 173 -

1st stage lever

2nd stage lever 8 9

12 6 4 3 pivot 20 19 13 21

7 5 2 1 PZT Strip 10 11 17 16

Y X Z

18 14 15

Fig. 9.5 Node information of the two-stage mechanism in the disk drive suspension.

Figure 9.5 shows the node information of the mechanism for SUGAR simulation. The netlist file of the geometry information is included in Appendix J with beam dimensions. Figure 9.6 (a) shows the effect of the second-stage pivot beam (nodes 9, 10 and nodes 15, 16, the stage close to output) width on the displacement amplification factor, with the pivot length fixed at 1000 m. The nomenclature used here is upstream classification. First-stage is the stage that is connected to the input. Second-stage is connected to the 1st-stage and the output. It is reversed from those used in previous Chapters for force-amplification leverage mechanisms. The amplification factor is almost constant at second-stage pivot beam width of less than 40 m, but with a maximum amplification factor of 32 at a pivot beam width of 30 m. Above second-stage pivot width of 40 m, the amplification factor decreases sharply with increasing pivot width. The final width of the second-stage pivot width is chosen to be 20 m.

- 174 -

The amplification factor as a function of the second-stage pivot length is shown in Figure 9.6 (b) for a fixed pivot width of 20 m. When the second-stage pivot length is about 300 m, the amplification factor drop dramatically. That is because the vertical displacement of the connection beam (nodes 7 and 8) almost approaches zero when the second-stage pivot length is close to the connection beam length of 400 m. With the second-stage pivot beam length greater than 300 m, the amplification factor increases with increasing pivot length and levels off at pivot length greater than 800 m. The final length of the second-stage pivot beam is chosen as 1000 m. The effect of first-stage (the stage close to input) pivot width on the amplification factor is shown in Fig. 9.7 (a) for a fixed pivot length of 1000 m. The amplification factor drops dramatically with increases in the pivot width up to 40 m. Then, further increases in the first-stage pivot does not cause any significant decreases in the amplification factor. In order to meet the specific device requirements, the final firststage pivot width of the mechanism is chosen to be 20 m. Figure 9.7 (b) shows the change in the amplification factor as the first-stage pivot length increases, with the pivot width fixed at 20 m. The amplification factor increases with increases in the first-stage pivot length. The final first-stage pivot length is chosen as 1000 m. All the beam dimensions are given in the netlist file in Appendix J. The entire two-stage leverage mechanism has a maximum amplification of 30 while the disk drive suspension with single-stage leverage mechanism has an amplification factor of 3. The two-stage leverage mechanism increases the amplification factor dramatically.

- 175 -

The gimbal needs to be robust enough to carry a vertical load at the tip, a 1-2 gram force along the Z-direction (normal to the structure). The second-stage pivot needs to be robust enough to sustain the bending moment generated by the Z-direction loading. The Z-direction bending spring constant should be 12-18 KN/m or greater [Chen, 2001]. The thick structure (75 micron) as well as the relatively wide flexure beam width (20 m or greater) ensure that the entire structure is within elastic range under typical loading. The dynamic response of the mechanism also needs to be considered for commercial applications [Chen, 2001]. For example, the resonant frequency of the mechanism

around Y direction should be greater than 10KHz while the mechanism carries a gimbal weight about 2 milligrams. Furthermore, the resonant frequency of the mechanism around X-direction should be greater than 2.5 KHz. These parameters were calculated through SUGAR simulations as shown in Fig. 9.8 for the Z-direction out-of-plane vibration mode and Fig. 9.9 for the Y-direction vibration mode. The resonant frequencies are 33KHz and 37KHz, respectively. A third-stage mechanism is not feasible due to design geometry constraints.

- 176 -

50

Displacement Amplification Factor

2nd Pivot Beam Length 1000 micron

40

30
Total A SUGAR

20

10

0 0 20 40 60 80 100

Width of 2nd Pivot, micron


Fig. 9.6 (a) The amplification factor as a function of the 2nd pivot width.

50

Displacement Amplification Factor

2nd Pivot Beam Width 20 micron

40

30

Total A SUGAR

20

10

0 0 200 400 600 800 1000 1200 1400 1600

Length of 2nd Pivot, micron


Fig. 9.6 (b) The amplification factor as a function of the 2nd pivot length.

- 177 -

60

Displacement Amplification Factor

1st Pivot Beam Length 1000 micron

50 40 30 20 10 0 0 20 40 60 80 100
Total A SUGAR

Width of 1st Pivot, micron


Fig. 9.7
60

(a) The effect of the 1st pivot width on the amplification factor.

Displacement Amplification Factor

1st Pivot Beam Width 20 micron

50 40 30 20 10 0 0 200 400 600 800 1000 1200 1400 1600

Total A SUGAR

Length of 1st Pivot, micron


Fig. 9.7 (b) The effect of the 1st pivot length on the amplification factor.

- 178 -

Fig. 9.8

Z-direction vibration mode displayed by SUGAR.

Fig. 9.9

Y-direction vibration mode displayed by SUGAR.

- 179 -

9.4 Application of Leverage Mechanisms for Displacement Amplification in a Silicon Microvalve


The design of a silicon microvalve for the proportional control of fluids is shown in Fig. 9.10(a) [Williams, 1999]. The input displacement is originated from a thermal expansion actuator (heatuator). A single-stage microleverage mechanism was used to amplify the input displacement generated by the thermal expansion actuator. The amplification factor is the lever ratio. When unactivated, the valve is open and fluid can flow down through inlet orifice to outlet port. Under activation, the output displacement at the tip of the lever arm (slider) reduces the orifice area and thus the flow. Figure 9.10(b) shows a proposed two-stage leverage mechanism to increase the displacement amplification factor. The two lever stages are both of the third-kind. The two-stage microleverage mechanism can greatly increase the amplification factor of the displacement. SUGAR simulation is used to study the output displacement changes with geometry changes of each component of the mechanism. A plot of the amplification factors (total amplification factor and the amplification of each lever-stage) as a function of the first-stage (close to input) pivot width is shown in Fig. 9.11(a). For a lever arm width of 40 m, the amplification factor does not change much when the width of the first-stage pivot increases. The amplification factor decreases noticeably with increasing of the first-stage pivot length, as shown in Fig. 9.11 (b). It is interesting to note that, for displacement amplification by a leverage mechanism, the effect of the flexure beam width and length on the amplification factor is reversed from that for the leverage mechanism for force amplification.
- 180 -

Figure 9.12 (a) shows the effect of connection beam width on the amplification factor. The amplification factor decreases significantly with increases in the connection beam width. The amplification factor increases slightly with the increase of the connection beam length, as shown in Fig. 9.12 (b). For the second-stage (the stage close to output) pivot, the amplification factor decreases with the increase of the pivot width and increases with the increase of the pivot length, as shown in Figure 9.13 (a) and (b), respectively. In summary, the flexure beams should be progressively thicker and shorter from the upstream (closer to output) to downstream lever stages, which is opposite to the case of leverage mechanisms for force amplification. For displacement amplification in a case of a microvalve with free-standing output, it is also important to investigate the possibilities of a three-stage leverage mechanism as schematically shown in Fig. 9.14 since the output side is freestanding and not connected to other structures. Figure 9.15 shows the amplification factor change with the change in the geometry of the first-stage pivot by SUGAR simulation of the three-stage mechanism shown in Fig. 9.4. The amplification factor increases with the increase of the first-pivot width, but decreases with the increase of the first-pivot length. Figure 9.16 shows the amplification factor change with the second-stage pivot geometry. It is found that the amplification factor increases slightly with increases of the second-stage pivot width up to 3 m, then decreases steadily with further increases in the pivot width. The effect of the second-stage pivot length on the total amplification factor is negligible, because the consequent increase in the amplification factor of the first-stage, A1, is balanced by the decrease in the amplification factor of the second-stage, A2. For the effect of the third- 181 -

stage pivot width and length change on the amplification factor, the SUGAR results are shown in Fig. 9.17. The amplification factor decreases sharply with the increase of the third-stage pivot width before leveling off at widths above 6 m; but increases rapidly with the increase of the third pivot length before leveling off at lengths above 160 m. With a three-stage leverage mechanism and free-standing output in this application case, the amplification factor can reach 500 or greater. By comparison, the second-stage leverage mechanism can have an amplification factor of 150, and the single-stage leverage mechanism amplification factor is 10.

- 182 -

Output displacement to open valve Z

Y Z

Fixed pushrod as pivot

Lever arm

Heatuator ro generate input displacement Movable pushrod as input beam

a.

Valve concept using single stage microlever Output displacement to open valve 1st stage pivot

Heatuator ro generate input displacement 2nd stage lever

1st stage lever 2nd stage pivot

b.

Valve concept using two-stage microlever

Fig. 9.10 (a) Schematic of a silicon microvalve with single-stage leverage mechanism (Williams 1999); (b) Two-stage microlever in the silicon microvalve.

- 183 -

160

25
Total A

140
Total A

20

120
A2 A1 (40) A1

15

100

10

80

1st Lever Arm Width = 40 & 20 micron 1st Pivot Length = 10 micron

60 2 4 6 8 10 12

0 14

Width of 1st Pivot, micron


Fig. 9.11 (a) Effect of first-stage pivot width change on the amplification factor.
160 25

140
Total A

20

120
A2

15

100
A1

10

80

60 0 20 40 60 80 100

0 120

Length of 1st Pivot, micron


Fig. 9.11 (b) Effect of first-stage pivot length change on the amplification factor.
- 184 -

Lever 1, 2 Disp. Amplification Factor

Displacement Amplification Factor

1st Lever Pivot Width = 2 micron Lever ratio: L1/l1=200/20, L2/l2=300/20

Lever 1, 2 Displ. Amplification Factor

Total Displ. Amplification Factor

180

25

150

Total A

20

120

A2

15

90

10

60

A1 5

30 0 2 4 6 8 10

0 12

Width of Connection Beam, micron


Fig. 9.12 (a) Effect of connection beam width change on the amplification factor.
160 25

140

Total A

20
A2

120
A1

15

100

10

80

60 0 20 40 60 80 100

0 120

Length of Connection Beam, micron


Fig. 9.12 (b) Effect of connection beam length change on the amplification factor.
- 185 -

Lever 1 &2 Disp. Amplification Factor

Total Disp. Amplification Factor

Connection Beam Width = 2 micron Lever ratio: L1/l1 = 10, L2/l2 = 15

Lever 1, 2 Disp. Amplification Factor

Total Displ. Amplification Factor

Connection Beam Length = 100 micron Lever ratio: L1/l1 = 200/20, L2/l2 = 300/20

180

25

150

Total A

20

120

A2

15

90

10

60

A1 5

30 0 2 4 6 8 10

0 12

Width of 2nd Lever Pivot Beam, micron


Fig. 9.13 (a) Effect of second-stage pivot width change on the amplification factor.
160 25

140

Total A

20
A2

120
A1

15

100

10

80

60 0 20 40 60 80 100

0 120

Length of 2nd Pivot Beam, micron


Fig. 9.13 (b) Effect of second-stage pivot width change on the amplification factor.
- 186 -

Lever 1 &2 Disp. Amplification Factor

Total Disp. Amplification Factor

2nd Pivot Beam Width = 2 micron Lever ratio: L1/l1 = 10, L2/l2 = 15

Lever 1, 2 Disp. Amplification Factor

Total Displ. Amplification Factor

2nd Pivot Beam Length = 100 micron Lever ratio: L1/l1 = 200/20, L2/l2 = 300/20

Output displacement to open valve 2nd stage lever 2nd stage pivot

1st stage lever 3rd stage lever 1st stage pivot

Heatuator ro generate input displacement

3rd stage pivot

Fig. 9.14

A three-stage leverage mechanism in the silicon microvalve.

- 187 -

600
1st, 2nd, 3rd Lever Ratio 10, 15, 20

30
Total A

500 400 300 200


A3

25 20 15 10 5
A1

A3

100 0 2 4 6 8 10

0 12

Width of 1st Pivot, micron


Figure 9.15 (a) Amplification factor change with the 1st pivot width

600
1st, 2nd, 3rd Lever Ratio 10, 15, 20

30 25 20 15
Total A

500 400 300 200


A2

A3

10 5
A1

100

0 0 10 20 30 40 50 60 70 80 90 100 110 120

Length of 1st Pivot, micron


Figure 9.15(b) Amplification factor change with the 1st pivot length

- 188 -

Lever 1, 2, 3 Displ. Ampli. Factor

Total Displ. Amplification Factor

Lever 1, 2, 3 Displ. Ampli. Factor

Total Displ. Amplification Factor

600
1st, 2nd, 3rd Lever Ratio 10, 15, 20

30 25 20 15 10
A2

500 400 300 200 100

Total A

A3

5
A1

0 0 2 4 6 8 10

0 12

Width of 2nd Pivot Beam, micron


Figure 9.16(a) Amplification factor change with 2nd pivot width

600
1st, 2nd, 3rd Lever Ratio 10, 15, 20

30 25 20 15 10
A2

500 400 300 200 100

Total A

A3

5
A1

0 0 20 40 60 80 100

0 120

Length of 2nd Pivot Beam, micron


Figure 9. 16(b) Amplification factor change with 2nd pivot length

- 189 -

Lever 1, 2, 3 Displ. Ampli. Factor

Total Displ. Amplification Factor

Lever 1, 2, 3 Displ. Ampli. Factor

Total Displ. Amplification Factor

900

30 25 20
A3

Total A

600

15
A2

300

10 5 0 12

A1

0 0 2 4 6 8 10

Width of 3rd Pivot Beam, micron


Figure 9.17(a) Amplification factor change with the 3rd pivot width

600

30 25 20 15
A2

500 400 300 200 100

A3

10 5

A1

0 0 40 80 120 160 200

0 240

Length of 3rd Pivot Beam, micron


Figure 9.17(b) Amplification factor change with 3rd pivot length

- 190 -

Lever 1, 2, 3 Displ. Ampli. Factor

Total Displ. Amplification Factor

1st, 2nd, 3rd Lever Ratio 10, 15, 20 Total A

Lever 1, 2, 3 Displ. Ampli. Factor

Total Displ. Amplification Factor

1st, 2nd, 3rd Lever Ratio 10, 15, 20

CHAPTER 10 CONCLUSIONS AND CONTRIBUTIONS

Recognizing the lack of systematic study on compliant microleverage mechanisms in the open literature and the fast-growing application of such mechanisms in MEMS, this thesis presents original work on a comprehensive analysis of compliant microleverage mechanisms, with the following key conclusions: 1. In Chapter 4, an extensive analytical model is developed for the single-stage compliant microleverage mechanism that forms the basic element. A general formulation of force amplification factor, A, of different kinds of single-stage microlever has been derived. The amplification factor depends not only on the ideal leverage ratio (L/l), the geometry of the pivot, but also the axial and bending spring constant of the output system. A too soft output system will decrease the amplification factor dramatically. 2. To gain insight into single-stage microleverage mechanism design, a term called the amplification coefficient, A*, has been defined as the product of (i) the sum of the bending spring constants at the pivot and output system and (ii) the sum of the reciprocals of the axial spring constants at the pivot and output system. The amplification factor increases as the amplification coefficient decreases. Among

- 191 -

the axial and bending spring constants of the pivot and output system, kvvp, kvvo,

k m p, and k m o, it is the larger rotational spring constant and smaller axial spring
constant that has more influence on the amplification coefficient and amplification factor. 3. The design and optimization of the single-stage microleverage mechanism in a resonant accelerometer is presented with the optimum geometry for the required compliance and flexibility. Both analytical and FEM approaches are used, together with the energy analysis. Good agreement is obtained between the results of second-order analytical modeling and those of FEM simulation with SUGAR. 4. In Chapter 5, a first-order analysis of the total force amplification factor, A, for a two-stage microleverage mechanism has been developed. The total amplification factor is approximated by the multiplication of the amplification factor of each microlever stage with the upstream microlever stage being treated as the output for the downstream microlever stage. 5. In order for a two-stage microlever to effectively amplify force, the compliance between the two microlevers must match as discussed in Section 5.3. More specifically, the axial spring constant of the microlever stage close to the output system needs to be in a specific region in order for the amplification factor of the microlever stage connected to the input system to be greater than 1. The connection and pivot beams in the microlever stage connected to the input system need to be as narrow and as long as possible and those in the microlever stage close to the output system should be wider and shorter in comparison.

- 192 -

6.

The input axial spring constant of a compound microlever is approximately the output axial spring constant reduced by the ideal amplification factor and the actual amplification factor of the compound microlever as seen in equation 5.13. The same reasoning applies to the microlever stage within a compound microlever.

7.

The increased resistance of a microlever to rotational or axial displacement when the output and pivot are at the different side of the lever arm leads to a lower force amplification factor. The optimum design of a two-stage microlever as analyzed in Section 5.4, should be the 1S/2S-1S/2S combinations, i.e., the pivot and connection beam of both the first- and second-stage microlever are on the same side of the lever arm.

8.

With multiple-stage (greater then 2) microlevers, the amplification factor has a multiplying effect and a general formulation of the total amplification factor is derived in Chapter 6. Also derived is the maximum amplification for a given output system and flexure beam dimension constraints set by micro-machining technology and available chip area. The compliance match theory also applies to the design of multi-stage microleverage mechanism. More specifically, the microlever stages in a compound microleverage mechanism for force amplification should increase in stiffness from input to output. In addition, a compound microlever should be designed with the minimum number of microlever stages to achieve the desired amplification factor.

- 193 -

9.

For the leverage mechanism design in a resonant output accelerometer, the optimum number of microlever stages is 2 to 3. The compliance match is hard to achieve when too many microlever stages are connected in series. A two-stage microlever with amplification of 80 was designed for the resonator accelerometer to amplify the inertial force and increase the sensitivity (Chapter 6 and 7). The device was fabricated with the SOI-MEMS process run. Testing of the device agree with the analytical results.

10.

Experimental verification of the design theory (i.e., analytical equations and SUGAR predictions) was carried out by using both micro-scale and macro-scale device models and described in Chapter 7 and 8. A 1S-2D type of two-stage microlever was fabricated by the SOI-MEMS process for inertial force amplification in the resonant accelerometer (Chapter 7). Through SUGAR simulation, the measurement sensitivity was shown to be greatly increased due to the force-amplification by the two-stage leverage mechanisms. A macro-scale model was made with aluminum strips as the flexure beams and aluminum tubes as the rigid lever arm. The experimental results agree qualitatively with the analytical equations and SUGAR predictions (Chapter 8).

11.

Leverage mechanism design for displacement amplification is analyzed with several amplifications, including in a disk drive suspension and a microvalve. Two-stage and multiple-stage of leverage mechanisms are designed and analyzed (Chapter 9).

- 194 -

12.

Microleverage mechanism design issues, such as the contradicting stiffness requirement, different trade-offs, maximum compressive stress and buckling, the effect of forces at different directions, are addressed in single- and multiple-stage microleverage mechanisms.

- 195 -

BIBLIOGRAPHY
1. Adams, S. G., Bertsch, F. M., Shaw, K. A., Hartwell, P. G., MacDonald, N. C., Moon, F. C., Capacitance Based Tunable Micromechanical Resonators, The 8th International Conference on Solid-State Sensors and Actuators-Transducers95, Stockholm, Sweden, 1995, vol.2. pp. 438-441. 2. Ananthasuresh, G. K., and Kota, S., Designing Compliant Mechanisms, Mechanical Engineering, v 117 n 11 November 1995, pp93-96. 3. Ananthasuresh, G. K., Frecker, M. I., Nishiwaki, S., Kikuchi, N., Kota, S., Topological Synthesis of Compliant Mechanism Using Multi-Criteria

Optimization, Journal of Mechanical Design, ASME Transactions, Vol.119, n 2 June 1997, pp238-245. 4. Ananthasuresh, G. K., Kota, S., Gianchandani, Y. A., A Methodical Approach to the Design of Compliant Micromechanisms. Technical Digest. Solid-State Sensor and Actuator Workshop, Hilton Head Island, SC, USA, 13-16 June 1994. Transducers 1994. pp. 189-92. 5. Bamford, R., Kuo, C.P., Glaser, R., Wada, B. k., Long Stroke Precision PZT Actuator, American Institute of Aeronautics and Astronautics, AIAA-95-1107CP, pp. 3278-3284. 1995. 6. Beckwith, T. G., Buck, N. L., Marangoni, R. D., Mechanical Measurement, third edition, Addison-Wesley Publishing Company.

- 196 -

7.

Bobbio, S. M., Kellam, M.D., Dudley, B.W., Goodwin-Johnson, S., Jones, S. K., Jacobson, J. D., Tranjan, F. M., and DuBois, T.D, Integrated Force Arrays, in Proc. IEEE MicroElectroMechanical Systems, Fort Lauderdale, FL, Feb. 7-10, 1993, pp.149-154.

8.

Boser, B. Capacitive Interfaces for Monolithic Integrated Sensors, Research meeting material. 1997.

9.

Brosnihan, T. J. et. al., Embedded Interconnect and Eectrical Isolation for Highaspect-ratio SOI Inertial Sensors, The 9th International Conference on SolidState Sensors and Actuators-Transducers, 97, Chicago, Illinois. pp.637-40, June 1997.

10.

Brosnihan, T. J., An SOI Based, Fully Integrated Fabrication Process for HighAspect-Ratio Microelectromechanical System, Ph.D. Dissertation, U. C. Berkeley, Berkeley, 1998.

11.

Chen, T. L., Horowitz R. Desing and Fabrication of PZT-Actuated Silicon Suspensions, Mechatronics 2000, Geogia Tech.

12. 13.

Chen, T. L. Personal Communication, 2000-2001. Cheshmehdoost, A., Jones, B. E., Design and Performance Characteristics of an Integrated High Capacity DETF-based Force Sensor, The 8th International Conference on Solid-State Sensors and Actuators-Transducers95, Eurosensors IX, pp. 389-410.

- 197 -

14.

Chiao, M. and Lin, L., Self-Buckling of Micromachined Beams Under Resistive Heating, J. of Microelectromechanical Systems, Vol. 9, No. 1, March 2000, pp.146-151.

15.

Clark, J. V., Zhou, N., and Pister, K. S. J., MEMS Simulation Using SUGAR v0.5, Solid-State Sensor and Actuator Workshop, Hilton-Head Island, South Carolina, June 8-11, 1998, pp191-196.

16.

Clark, J. V., Zhou, N., and Pister, K. S. J., Modified Nodal Analysis for MEMS with Multi-Energy Domains, International Conference on Modeling and Simulation of Microsystems, Semiconductors, Sensors and Actuators, San Diego, CA, March 27-29, 2000,

17.

Cohn, M. B., Assembly Techniques for Microelectromechanical Systems, Ph.D. Dissertation, U. C. Berkeley, Berkeley, 1997.

18.

Fedder, G. K., Simulation of Microelectromechanical Systems, Doctoral Dissertation, U. C. Berkeley, Berkeley, 1994.

19.

Frecker, M. I., Kota, S., Kikuchi, N., Optical Design of Compliant Mechanisms for Smart Structures Applications". Proceedings of the SPIE- The International Society for Optical Engineering, vol. 3323, (Smart Structures and Materials 1998: Mathematics and Control in Smart Structures, San Diego, CA) SPIE-Int. Soc. Opt. Engr, 1998. P. 234-42.

20.

Hall. Jr., A. S., Mechanisms and Their Classification, Transactions of the First Conference on Mechanisms, Machine Design, December 1953, pp.174-180.

- 198 -

21.

Harris, C. M., Crede, C. E., Shock and Vibration Handbook 2nd Ed., McGrawHill, New York, 1976.

22.

Hartog, J. P. Den, Mechanical Vibrations, 4th Edition, McGraw-Hill book Company, Inc. 1856

23.

Her, I., Chang, J. C. A linear Scheme for the Displacement Analysis of Micropositioning Stages with Flexure Hinges, Machine Element and Machine Dynamics, ASME 1994, DE-Vol 71, pp. 517-525.

24.

Hibbitt, K. S., Inc. 1080 Main Street, Pawtucket, Rhode Island. ABAQUS Users Manual, 5.4 ed., 1994.

25.

Ho, C. H., Tai, Y. C., Micro-Electro-Mechanical-System (MEMS) and Fluid Flows, Annu. Rev. Fluid Mech. 1998. 30: pp579-612.

26.

Howe, R. T., 1987, Resonant Microsensors, Fourth International Conference on Solid-State Sensors and Actuators, Tokyo, Japan, pp.843-848.

27.

Howe, R. T., 1984, Integrated Silicon Electromechanical Vapor Sensors, Ph. D Dissertation, U. C. Berkeley, Berkeley, 1994.

28.

Jenson, B. D., Howell, L. L., Gunyan, D. B., Salmon, L. G., The Design and Analysis of Compliant MEMS Using the Pseudo-Rigid-Body Model, DSC-Vol. 62/HTD-Vol. 354, Microelectromechanical System (MEMS) ASME 1997.

29.

Keller, C. G. and Howe, R. T., Nickel-Filled HEXSIL Thermally Actuated Tweezers, The 8th International Conference on Solid-State Sensors and Actuators-Transducers95, Stockholm, Sweden, 1995.

- 199 -

30.

Keller, C. G., Howe, R. T., Hexsil tweezers for teleoperated microassembly, in Proc. IEEE MEMS Workshop, 1997, pp.72-77.

31.

Keller, C. G., Ferrari, M., Milti-scale Polysilicon Structures,1994 Solid-State Sensor and Actuator Workshop, Hilton Head Is., SC, June, 1994, pp132-137.

32.

Kikuchi, N., Nishiwaki, S., Fonseca, J. S., Silva, E. C., Design Optimization Method for Compliant mechanisms and material Microstructure, Computer Methods in Applied Mechanics and Engineering, V 151, n3-4, Jan. 20 1998, pp.401-417.

33.

Kota, S., Ananthasuresh, G. K., Crary, S. B., Wise, K.D., Design and Fabrication of Microelectromechanical Systems, Journal of Mechanical Design, December 1994, Vol. 116, pp1081-1088.

34.

Larsen, U. D., Sigmund, O., and Bouwstra, S., 1997. Design and Fabrication of Compliant Micromechanisms and Structures with Negative Poissons Ratio. Journal of Microelectromechanical Systems, Vol. 6, No. 2, June,1997, pp.99-106.

35.

Laximarayana Saggere and Kota, S., Synthesis of Distributed Compliant Mechanisms For Adaptive Structures Application: An Elasto-kinematic Approach, ASME Design Engineering Technical Conference, Sept. 1997.

36.

Lee, K. B., Cho, Y. H., A Triangular Electrostatic Comb Array for Micromechanical Resonant Frequency Tuning, Sensors and Actuators A 70 (1998), pp. 112-117.

- 200 -

37.

Lee, K. B., Cho, Y. H., Frequency Tuning of a Laterally Driven Microresonator Using an Electrostatic Comb Array of Linearly Varied Length, The 9th International Conference on Solid-State Sensors and Actuators- Transducers 97, Chicago. Illinois, pp. 113-116. June 1997.

38.

Lin, L., Howe, R. T., and Pisno, A. P., A Passive, In Situ Micro Strain Gauge, in Proc. IEEE MicroElectroMechanical Systems, Fort Lauderdale, FL, 1993, pp. 201-206.

39.

Mach, E., The Science of Mechanics - A Critical and History Account of Its Development, ( Open Court, LA Salle, IL, 1960).

40.

Main, J. A., and Garcia, E. and Newton, D. V., 1995. Precision Position Control of Piezoelectric Actuators Using Charge Feedback. Journal of Guidance, Control, and Dynamics, Vol. 18, No.5, September-October, 1995, pp1068-1073.

41.

Markus, K. W. and Koester, D. S., Multi-user MEMS process (MUMPs) introduction and design rules, MCNC Electron., Oct. 1994.

42.

Midha, A., Norton, T. W., Howell, L. L,

On the Nomenclature and

Classification of Compliant Mechanism: The Component of Mechanisms, Flexible Mechanisms, Dynamics, and Analysis, ASME 1992, DE-Vol. 47, pp.237245. 43. Nguyen, C. T. - C., Micromechanical Signal Processors, Ph. D. dissertation, Dept. Elec. Eng. Comput. Sci., Univ. California-Berkeley, Dec. 1994.

- 201 -

44.

Nishiwaki, S., Min S., Ejima, S., Kikuchi, N., Structural Optimization Considering Flexibility, JSME, International Journal, Series C. v 41, n3, Sep. 1998, pp.476-484.

45.

Nishiwaki, S., Frecker, M. I., Min, S., Kikuchi, N., Topology Optimization of Compliant Mechanisms Using the Homogenization Method, International Journal for Numerical Methods in Engineering 42. P. 535-559 (1998).

46.

Ohanian, H. C., Rotational Motion and the Law of the Lever. American Journal of Physics, Vol. 59, ( no.2), Feb. 1991. Pp182.

47.

Paros, J. M. and Weisbord, L., 1965, How to Design Flexure Hinges, Machine Design, Vol. 37 n. 27, pp.151-156.

48.

Petersen, K. E., Silicon as a Mechanical Material Proc. IEEE, vol. 70, no. 5, pp. 420-457, May 1982.

49.

Putty, M. W., Polysilicon Resonant Microstructures, M. S. Thesis, University of Michigan, 1988.

50. 51.

Rao, S. S., 1990, Mechanical Vibrations, 2nd Ed., Addison-Wesley, Reading, MA. Roessig, T. A., Howe, R. T., Pisano, A. P., Smith, J. H., Surface-Micromachined Resonant Accelerometer, The 9th International Conference on Solid-State Sensors and Actuators- Transducers 97, Chicago. Illinois, pp. 859-862.

52.

Roessig, T. A., Howe, R. T., Pisano, A. P., Smith, J. H., SurfaceMicrormachined Resonant Force Sensor, Proceedings of the ASME Dynamic

- 202 -

System and Control Division, DSC-Vol. 57-2, 1995 ASME Francisco,1995,pp.871-876. 53.

IMECE, San

Roessig, T. A., Surface Micromachined Resonant Accelerometer, Ph.D. Dissertation, U. C. Berkeley, Berkeley, 1998.

54.

Roessig, T. A., Surface Micromachined Resonant Force Transducers, M-S Report, U. C. Berkeley, 1995.

55.

Saggere, L., Kota, S., Design of Adaptive Structures Using Compliant Mechanisms, Proceedings of the SPIE-The International Society for Optical Engineering, vol. 3329, pt1-2, (Smart Structures and Materials 1998: Smart Structures and Integrated System, San Diego, CA, 2-5 March 1998.) SPIE Int. Soc. Opt. Eng, 1998. P.672-6.

56.

Salamon, B. A., Midha, A., 1998. An Introduction to Mechanical Advantage in Compliant Mechanisms. Journal of Mechanical Design, Vol 120, No. 2, June,1998, p311-315.

57.

Samuelson, M., Garcia, E., Otto, J., Paine, J., and Main, J., 1997, Actuator Systems for Precision Motion Control. Paper Number 97-063.

58.

Shigley, J. E., Mitchell, L. D., Mechanical Engineering Design, McGraw-Hill Book Company.

59.

Sigmund, O. On the Design of Compliant Mechanisms Using Topology Optimization, Mechanics of Structures and Machines, V 25, n4, Nov. 1997, pp.493-524.

- 203 -

60. 61.

Singer, F. L., Strength of Materials, 2nd Edition, Happer & Row Publishers. Sniegowski, J. J., and Smith, C., An Application of Mechanical Leverage to Microactuation, The 8th International Conference on Solid-State Sensors and Actuators-Transducers95, Stockholm, Sweden, 1995, 325 PB 9.

62.

Su, X-P. S., and Yang, H. S., 2000 (1), Single-Stage Microleverage Mechanism Optimization in a Resonant Accelerometer accepted to publish in Structure and Multidisciplinary Optimization.

63.

Su, X-P. S., and Yang, H. S., 2000 (2), Two-Stage Microleverage Mechanism Optimization in a Resonant Accelerometer accepted to publish in Structure and Multidisciplinary Optimization.

64.

Su, X-P. S., and Yang, H. S., 2000 (3), Design of Compliant Microleverage Mechanism Sensors and Actuators, January, 2001, vol. 87 issue 3 pp.146-156.

65.

Su, X-P. S., and Yang, H. S., 2000 (4), Analytical Modeling and FEM Simulation of Single-Stage Microleverage Mechanism International Journal of Mechanical Sciences, accepted for publish.

66.

Su, X-P. S., Alice M. Agogino and Yang, H. S., 2000, Two-Stage Compliant Leverage Mechanism Design and Application, accepted for publishing in Journal of Mechanical Design.

67.

Takeshima, N., Gabriel, K. J., Ozaki, M., Takahashi, J., Horiguchi, H., and Fujita, H., Electrostatic Parallelogram Actuators, in Proc. 1991 International

- 204 -

Conference on Solid-State Sensor and Actuators, San Francisco, June 24-27, 1991, Transducer 91, pp.63-66. 68. Tang, W. C., Electrostatic Comb Drive for Resonant Sensor and Actuator Application, Ph.D. Dissertation, U. C. Berkeley, 1990. 69. Tang, W. C., Nuyen, T.-C. H., and Howe, R. T., Laterally Driven Polysilicon Resonant Microstructures, Sensors and Actuators, 20, pp.25-32 (1989) 70. Tang, W. C., Nuyen, T.-C. H., Judy, M. W. and Howe, R. T., Electrostatic-Comb Drive of Lateral Polysilicon Resonators, Sensors and Actuators A, A21- 23, pp.328-331 (1990) 71. Tang, W. C., Lim, M. G., Howe, R. T., Electrostatic comb drive levitation and control methods, Journal of Microelectromechanical Systems, pp.170-178 (1992). 72. Timoshenko, S., Young, D. H., and Weaver, W., Jr., 1974, Vibration Problems in Engineering, 4th Ed., John Wiley and Sons, New York 73. Timoshenko, S. P. and Goodier, J. N., Theory of Elasticity, 3rd ed. New York: McGraw-Hill, 1970. 74. Timoshenko, S., Young, D. H., Elements of Strength of Materials, Fifth Edition, D. Van Nostrand Company. 75. Weigold, J. W., Wong, A. C. Nguyen, C. T.-C., Pang, S. W., A Merged Process for thick Single-Crystal Si Resonators and BiCMOS Circuitry, Journal of Microelectromechanical Systems, Vol. 8. No. 3. September 1999.

- 205 -

76.

Williams, K., A Silicon Microvalve for the Proportional Control of Fluids, Transducers 99, pp.1804-1807.

77.

Zhou, N., Clark, J.V. and Pister, K.S.J., Nodal Simulation for MEMS Design Using SUGAR v0.5, 1998, International Conference on Modeling and Simulation of Microsystems Semiconductors, Sensors and Actuators, Santa Clara, CA, April 6-8, 1998. pp.308-313.

- 206 -

APPENDICES
APPENDIX A ABAQUS Input File For Single Lever:

*HEADING CANTILEVER BEAM *NODE, NSET=BOTTOM 1,0.0,0.0 211,210.0,0.0 *NGEN, NSET=BOTTOM 1,211,1 *NODE, NSET=TOP 5001,0.0,20.0 5211,210.0,20.0 *NGEN, NSET=TOP 5001,5211,1 *NFILL,NSET=INSIDE BOTTOM,TOP,5,1000 *ELEMENT,ELSET=BULK,TYPE=CPS4R 1,1,2,1002,1001 *ELGEN,ELSET=BULK 1,210,1,1,5,1000,210 *SOLID SECTION,ELSET=BULK,MATERIAL=POLI 2 *MATERIAL, NAME=POLI

- 207 -

*ELASTIC 1.5E11,0.3 *DENSITY 2330.0 *NSET,NSET=BB1 5001,5002,5003 *NODE,NSET=BTOP1 11001,0.0,26.0 11002,1.0,26.0 11003,2.0,26.0 *NSET,NSET=BB2 5011,5012,5013 *NODE,NSET=BTOP2 29011,10.0,44.0 29012,11.0,44.0 29013,12.0,44.0 *NFILL,NSET=BIN1 BB1,BTOP1,6,1000 *NFILL,NSET=BIN2 BB2,BTOP2,24,1000 *ELEMENT,ELSET=BEAM1,TYPE=CPS4R 2001,5001,5002,6002,6001 *ELEMENT,ELSET=BEAM2,TYPE=CPS4R 2003,5011,5012,6012,6011 *ELGEN,ELSET=BEAM1 2001,2,1,1,6,1000,4 *ELGEN,ELSET=BEAM2 2003,2,1,1,24,1000,4 *SOLID SECTION,ELSET=BEAM1,MATERIAL=BEAM1 2

- 208 -

*MATERIAL, NAME=BEAM1 *ELASTIC 1.5E11,0.3 *DENSITY 2330 *SOLID SECTION,ELSET=BEAM2,MATERIAL=BEAM2 2 *MATERIAL, NAME=BEAM2 B *ELASTIC 1.5E11,0.3 *DENSITY 2330 *BOUNDARY BTOP1,1,2 BTOP2,1,2 *STEP, PERTURBATION *STATIC *CLOAD 211, 2, +1.25E0 *EL PRINT, POSITION=AVERAGED AT NODES, SUMMARY=YES S11, E11 SF *RESTART, WRITE *END STEP

- 209 -

APPENDIX B Netlist Files for Pivot Design


I. Vertical Pivot
a b b b b a f 2 2 3 4 4 6 5 1 3 4 5 6 7 1 8e-6 6e-6 20e-6 200e-6 200e-6 8e-6 1.5e-9 -90 +90 0 0 -90 -90 90 16e-6 2e-6 20e-6 20e-6 2e-6 16e-6 0

II. Horizontal Pivot


a b b b b a f 2 2 3 4 4 6 5 1 3 4 5 6 7 1 8e-6 6e-6 20e-6 200e-6 200e-6 8e-6 1.5e-9 0 0 0 0 -90 -90 90 16e-6 2e-6 20e-6 20e-6 2e-6 16e-6 0

III. Combine Pivot


a b b b b b a f 1 1 2 3 4 4 6 5 11 2 3 4 5 6 7 1 8e-6 6e-6 6e-6 20e-6 200e-6 200e-6 8e-6 1.5e-9 90 0 -90 0 0 -90 -90 90 16e-6 2e-6 2e-6 20e-6 20e-6 2e-6 16e-6 0

- 210 -

APPENDIX C Mathematica Files For Analytical Analysis


I. Single-Stage Microleverage Mechanism:
In[6]:= Solve[{Kvf (L1 T - D) - Kvp D == Fin, Kvf L1^2 T + (Ktp + Ktf) T - Kvf D L1 == Fin L2}, {T, D}] Out[6]= {{D -> Fin/(-Kvf Kvp) + (Kvf L1 (-(Fin Kvf L1) Fin (-Kvf - Kvp) L2)) ((-Kvf - Kvp)/ (-(Ktf Kvf) - Ktp Kvf - Ktf Kvp - Ktp Kvp - Kvf Kvp L1^2 )),T -> -((-(Fin Kvf L1) - Fin (-Kvf - Kvp) L2) / (-(Ktf Kvf)- Ktp Kvf - Ktf Kvp - Ktp Kvp - Kvf Kvp L1^2 ))}}

L1 = 0.00001; L2 = 0.00021;Fin = 0.0000000015; Ktp = 0.000000033;Kvp = 100000; Kvf = 25000; Ktf = 0.0000000083;

NSolve[{Kvf (L1 T - D) - Kvp D == Fin, Kvf L1^2 T + (Ktp + Ktf) T - Kvf DL1 == Fin L2}, {T, D}]

Out[5]= {{T -> 1.52844 10E-7 , D -> 2.93688 10E-13 }} m = Kvf (L1 T - d)/Fin Out[10]= 20.1667

In[12]:= K=Fin/(L2 T - d) Out[12]= 48.0615 ***************************************************************

II. Second-Stage Microleverage Mechanism:


L1 = 0.00001; L2 = 0.00021;Fin = 0.0000000015; Kvp = 100000; Ktp = 0.000000033; Kvf = 48.0615; Ktf = Ktp = 0.000000033;

NSolve[{Kvf (L1 T - De) - Kvp De == Fin, Kvf L1^2 T + (Ktp + Ktf) T - Kvf DeL1 == Fin L2}, {T, D}] Out[4]= {{T -> 4.44881 10E-6 , De -> 6.37858 10E-15 }} t = 0.00000444; De = 0.00000000000000638;

- 211 -

In[9]:= m = Kvf (L1 t - De)/Fin Out[9]= 1.42242 In[10]K=Fin/(L2 t - De) Out[10]= 1.60876 *****************************************************************

III. Third-Stage Microleverage Mechanism:


L1 = 0.00001; L2 = 0.00021;Fin = 0.0000000015; Kvp = 100000; Ktp = 0.000000033; Kvf = 1.60876; Ktf = 0.000000033;

NSolve[{Kvf (L1 T - De) - Kvp De == Fin, Kvf L1^2 T + (Ktp + Ktf) T - Kvf DeL1 == Fin L2}, {T, De}] Out[4]= {{T -> 4.76112 10E-6 , De -> -1.42338 10E-14 }} In[5]:= t = 0.0000047; d = -0.000000000000014; In[7]:= m = Kvf (L1 t - d)/Fin Out[7]= 0.0504228

- 212 -

APPENDIX D Netlist Files For Resonant Accelerometer With Single-Stage Microleverage Mechanism
A b b b b b b b a b b b b b a b f f 2 2 3 4 5 6 7 8 9 7 5 12 13 14 15 13 11 17 1 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 1 1 8e-6 4e-6 200e-6 20e-6 20e-6 24e-6 10e-6 6e-6 4e-6 200e-6 20e-6 24e-6 10e-6 6e-6 4e-6 200e-6 1.5e-9 1.5e-9 90 -90 -90 -90 0 -90 180 90 90 0 180 -90 0 90 90 180 90 90 16e-6 12e-6 2e-6 12e-6 20e-6 2e-6 20e-6 2e-6 8e-6 20e-6 20e-6 2e-6 20e-6 2e-6 8e-6 20e-6 0 0

- 213 -

APPENDIX E Netlist File for Resonant Accelerometer with Two-Stage Microleverage Mechanism
a b b b b b b b a b b b b b a b b b b b a b b b b a f f 2 2 3 4 5 6 7 8 1 3 4 5 6 7 8 9 8e-6 4e-6 100e-6 20e-6 20e-6 6e-6 10e-6 6e-6 4e-6 200e-6 20e-6 6e-6 10e-6 6e-6 4e-6 200e-6 60e-6 10e-6 200e-6 100e-6 8e-6 60e-6 10e-6 200e-6 100e-6 8e-6 1.5e-9 1.5e-9 90 -90 -90 -90 0 -90 180 90 90 0 180 -90 0 90 90 180 -90 180 180 -90 -90 -90 0 0 -90 -90 90 90 16e-6 12e-6 8e-6 12e-6 20e-6 4.0e-6 40e-6 4e-6 8e-6 40e-6 20e-6 4e-6 40e-6 4e-6 8e-6 40e-6 2e-6 20e-6 20e-6 2e-6 16e-6 2e-6 20e-6 20e-6 2e-6 16e-6 0 0

9 10 7 11 5 12 12 13 13 14 14 15 15 16 13 17 11 25 25 24 24 23 24 26 26 27 17 18 18 19 19 20 19 21 21 22 20 1 23 1

- 214 -

APPENDIX F Netlist File for Calculating Resonator Frequency

1. DETF with Comb-drive as Center Attached mass


a b b b b b a 2 2 3 3 6 6 4 1 3 4 6 7 8 5 8e-6 100e-6 100e-6 20e-6 60e-6 60e-6 8e-6 90 -90 -90 0 90 -90 -90 16e-6 2e-6 2e-6 4e-6 12e-6 12e-6 16e-6

2. DETF with Capacitive Tuning


a b b b b b b b b b a g g 2 2 21 31 3 32 41 3 6 6 4 21 32 1 21 31 3 32 41 4 6 7 8 5 31 23 41 34 8e-6 30e-6 60e-6 10e-6 10e-6 60e-6 30e-6 20e-6 60e-6 60e-6 8e-6 90 -90 -90 -90 -90 -90 -90 0 90 -90 -90 16e-6 2e-6 2e-6 2e-6 2e-6 2e-6 2e-6 4e-6 12e-6 12e-6 16e-6

33 60e-6 44 60e-6

-90 1e-7 4e-6 4e-6 0 -90 1e-7 4e-6 4e-6 0

- 215 -

APPENDIX G Netlist File for Simulation of the DETF Resonator


a1 a2demf p1 [2 b1 b2dem b2 b2dem p1 [2 p1 [4 a2 a2demf p1 [5 b1c b2dem p1 [3 b2c b2dem p1 [3 b3c b2dem p1 [4 b4c b2dem p1 [7 b5c b2dem p1 [8 b6c b2dem p1 [9 1] 3] 6] 5] 4] 7] 8] 9] 9] [l=5e-6 [l=5e-6 [l=10e-6 [l=20e-6 [l=20e-6 [l=5e-6 [l=5e-6 oz=90 oz=-90 oz=-90 oz=0 oz=0 oz=-90 oz=90 oz=0 oz=0 oz=90 oz=-90 oz=0 oz=0 w=10e-6 w=2e-6 w=10e-6 w=2e-6 w=2e-6 w=2e-6 w=2e-6 w=2e-6 w=2e-6 w=2e-6 w=2e-6 w=12e-6 w=12e-6 R=100] R=1000] R=100] R=1000] R=500] R=500] R=500] R=1000] R=1000] R=100] R=100] R=100] R=100]

[l=100e-6 oz=-90 [l=100e-6 oz=-90

10] [l=2e-6

b2m b2dem p1 [10 15] [l=2e-6 b3m b2dem p1 [10 16] [l=60e-6 b4m b2de c1 c2 p1 [10 17] [l=60e-6 p1 p1 [15 11][l=4e-6

comb2d comb2d

N=18 gap=3e-6] w=10e-6 w=10e-6 R=100] R=100] N=18 gap=3e-6]

a3 a2demf a4 a2demf

p1 [11 12] [l=5e-6

[9 13] [l=4e-6 oz=180

p1 [13 14] [l=5e-6 oz=180

v1 volc v2 volc v3 volc v4 volc

* * * *

[15 20] [V=10 sv=5] [9 20] [V=10 sv=5] [11 20] [V=0] [13 20] [V=0] * [20] []

e1 eground

- 216 -

APPENDIX H Netlist File for Simulation of the 1S-2D type Two-stage Leverage Mechanism in the Macro-model
a b b b b b b b a b b b b b a b b b b b a b b b b a f f 2 2 3 4 5 6 7 8 9 8 5 12 13 14 15 14 11 25 24 24 26 17 18 19 19 21 20 23 1 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 25 24 23 26 27 18 19 20 21 22 1 1 8e-3 4e-3 85e-3 20e-3 20e-3 22e-3 23e-3 22e-3 4e-3 221e-3 20e-3 22e-3 23e-3 22e-3 4e-3 221e-3 85e-3 19e-3 257e-3 136e-3 8e-3 85e-3 19e-3 257e-3 136e-3 8e-3 1.5e-3 1.5e-3 90 -90 -90 -90 0 -90 0 90 90 0 180 -90 180 90 90 180 -90 0 180 -90 -90 -90 180 0 -90 -90 90 90 16e-3 20e-3 3.6e-3 20e-3 20e-3 1.8e-3 40e-3 1.8e-3 8e-3 40e-3 20e-3 1.8e-3 40e-3 1.8e-3 8e-3 40e-3 2e-3 20e-3 20e-3 2e-3 16e-3 2e-3 20e-3 20e-3 2e-3 16e-3 0 0

- 217 -

APPENDIX I Netlist File for Simulation of the 1S-2S type Two-stage Leverage Mechanism in the Macro-model
a b b b b b b b a b b b b b a b b b b b a b b b b a f f 2 2 3 4 5 6 7 8 9 8 5 12 13 14 15 14 11 25 24 24 26 17 18 19 19 21 20 23 1 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 25 24 23 26 27 18 19 20 21 22 1 1 8e-3 4e-3 85e-3 20e-3 20e-3 22e-3 23e-3 22e-3 4e-3 221e-3 20e-3 22e-3 23e-3 22e-3 4e-3 221e-3 85e-3 19e-3 257e-3 136e-3 8e-3 85e-3 19e-3 257e-3 136e-3 8e-3 1.5e-3 1.5e-3 90 -90 -90 -90 0 -90 0 90 90 0 180 -90 180 90 90 180 -90 0 180 90 -90 -90 180 0 90 -90 90 90 16e-3 20e-3 3.6e-3 20e-3 20e-3 1.8e-3 40e-3 1.8e-3 8e-3 40e-3 20e-3 1.8e-3 40e-3 1.8e-3 8e-3 40e-3 2e-3 20e-3 20e-3 2e-3 16e-3 2e-3 20e-3 20e-3 2e-3 16e-3 0 0

- 218 -

APPENDIX J Netlist File for Simulation of the Two-stage Leverage Mechanism in the Disk-drive suspension
a b b b b b b b b a b b b b b b b b a b b f f 2 2 3 4 4 6 7 8 9 10 8 3 20 20 19 18 14 15 16 14 12 5 21 1 3 4 5 6 7 8 9 10 11 12 20 21 19 18 14 15 16 17 13 13 1 1 200e-6 1000e-6 100e-6 400e-6 500e-6 4000e-6 400e-6 500e-6 1000e-6 200e-6 5500e-6 100e-6 400e-6 500e-6 4000e-6 400e-6 500e-6 1000e-6 200e-6 5500e-6 2000e-6 1.5e-4 1.5e-4 0 180 90 0 90 0 90 0 -90 -90 180 -90 0 -90 0 -90 0 90 90 180 -90 180 0 1600e-6 20e-6 100e-6 20e-6 100e-6 100e-6 20e-6 100e-6 20e-6 1600e-6 100e-6 100e-6 20e-6 100e-6 100e-6 20e-6 100e-6 20e-6 1600e-6 100e-6 1000e-6 0 0

- 219 -

You might also like